Entanglement re-examined: Since Bell got it wrong, maybe Einstein had it right

Antony R. Crofts

Department of Biochemistry and Center for and Quantitative Biology University of Illinois at Urbana-Champaign, Urbana IL 61801

Correspondence: A.R. Crofts Department of Biochemistry University of Illinois at Urbana-Champaign 417 Roger Adams Lab 600 S. Mathews Ave Urbana, IL 61801 Phone: (217) 333-2043 Email: [email protected]

1 Bell’s treatment of the entanglement question and the outcome of subsequent experi- ments to test his hypothesis are generally taken as demonstrating that at the level nature is non-local (NL). From standard quantum mechanical (QM) predicates, when coincidences are counted as a function of settings on measurement at separate stations, tests using pho- ton pairs are expected to show the same amplitude (full “visibility”) sinusoidal curves independ- ent of orientation of the frame. This NL behavior, claimed to demonstrate non-locality, cannot be expected from objective local (OL) treatments. Experimental confirmation of this dif- ference is presented as one of the major achievements of 20th century physics because it con- strains all subsequent theoretical discussion of process at this level to a non-local perspective. In this paper, I argue that this edifice is built on sand. Einstein, Podolsky and Rosen (EPR) started the entanglement debate by framing it in terms of a model, necessarily local, in which quantum entities have intrinsic properties (“elements of reality”). When Bell joined the discussion, the OL model he proposed precluded consideration of vectorial properties, and then limited the behavior expected, and could not match the NL expectations. In Part A of the paper, I explore through discrete math simulation, a vectorial model (vOL) following Einstein’s expecta- tions. The simulation asks what would happen if with defined were used in typical experimental tests, and shows that when the photon frame is aligned with the reference polarizer the outcome is the same as in the NL model. This indicates that the inequalities Bell used to exclude local realism were in error. In analysis of this discrepancy, it is clear that Bell, in excluding vectorial properties, chose an unrealistic OL model. Experimental tests failed to find the observed outcome not because nature is necessarily non-local, but because the model, and justifications for a limit of ≤2 based on it, were wrong. Nevertheless, one feature expected from the NL treatment, the full-visibility on rotation of frames, is not explained by the vOL model. In Part B of the paper, I ask how well-justified the NL predicates are, and how plausible are extensions proposed to overcome tensions with classical constraints. Three predicates from QM lead to the NL expectation: (i) that the requires evolution to a measure- ment context in a superposition of states; (ii) that the pairs are in a dichotomy of H and V states that determine their behavior at discriminators; (iii) that the matrix operations describes a real process through which pairs are actualized with vectors aligned in frame with the reference polarizer. In the first QM predicate, quantum uncertainties preclude assignment of vectorial properties to quantum entities, demanding treatment through a superposition of states. However,

2 I suggest that this is a poor starting point because the uncertainty argument, originally applied to conjugate properties of a single electron, has no logical justification when applied to the two photons of a pair, interrogated through refraction and detected separately. The ontic dichotomy of spin states of the second predicate has no basis in physics, and the spin states determine be- havior at the by assignment of vectorial agency as propensities. The assignment of a vectorial role is also unjustified. In the third predicate, by representing the propensities in binary form in the matrices, the operations implement an unnatural alignment of photon and polarizer frames. This creates a Maxwellian demon and yields outcomes that are in contravention with second law and relativistic constraints. Given ambiguities in description of protocols, the simula- tion shows that results reported may have a simpler interpretation; manipulations in state prepa- ration lead to alignment. Rather than supporting non-locality, the results suggest local realistic explanations are more plausible than the conventional NL ones. It has been suggested that difficulties with the NL treatment can be ameliorated by later theoretical developments in the wavy realm. Dirac developed (QED) by co-opting Maxwell’s electromagnetic treatment to the interaction of photons with elec- trons through their fields, with the potential for action at a distance. Some theoretical sophistica- tion is required in applying the wavy treatment to process at the quantum level because photons are neutral. Feynman’s path integral approach extended the approach by considering each path as separately quantized. In effect, Maxwell’s approach could be used to calculate phase delays aris- ing from differences in path length, giving interference effects modulating probabilities for pho- tons travelling different paths. However, quantized interactions necessarily involve processes at that level, and neutral entities cannot act through fields. A more natural approach is to treat pho- tonic interactions in terms of momentum transfer. Work can be applied either in generation of a photon by displacement of an electron, or in displacement of an electron by a photon on interac- tion by absorption or refraction. As in Maxwell’s treatment, both generation and interaction are constrained by the two properties that can be measured, frequency and orientation. A full treat- ment based on a quantized transfer of momentum has previously been suggested. In effect, this momentum-based scenario is that implemented in my simulation. Maxwell’s electromagnetic perspective underlies much of present theoretical discussion, but transfer of momentum makes for a much simpler treatment, and brings new insights into processes on scales from the quantum to the cosmic, which might provide a more promising path forward.

3 INTRODUCTION The entanglement debate started when Einstein, Podolsky and Rosen1 introduced the so- called “EPR-paradox” (Einstein’s most cited paper2). This crystallized earlier arguments3-7 with Bohr over the foundations of (QM), and focused the debate through consideration of how correlations between two quantum objects should be represented when they are generated through a common transition. The EPR model was of a pair of entities, correlated through a common cause, which then separated to space-like distance before measurement. EPR pointed out that if the correlated properties were intrinsic to each entity (“elements of reality”), local measurement of one would allow the observer to infer a property of the other, allowing direct determination of pairwise correlations in a population. The predicates of the Copenhagen interpretation precluded such a pic- ture8,9. Measurement at this scale introduces uncertainty arising from quantized energy exchanges, accommodated by requiring an indeterminate superposition of all possible states of entangled pairs in the evolution to separate measurement stations for analysis. In the common wavefunction describing this non-local (NL) state, partners are correlated through a dichotomy of their spin states (the phase difference), but vectorial properties are indeterminate. Determinate entities with real vectors are actu- alized on operation of the Pauli matrices, which aligns the photon (or electron) frame with the dis- criminator frame. Measurement of an object at one location then appears to cause simultaneous actu- alization of a partner with complementary properties at a distant other. This would require transfer of energy and/or information faster than light, - that “spooky action at a distance” Einstein so disliked10. Some thirty years later, in “probably the most celebrated result in the whole of twentieth-cen- tury physics”11,12, Bell13-19 analyzed a population of electron pairs generated with opposite spin9. He compared the expectations from an objective local (OL) model with those from an NL model of the same population in superposition and showed them to be different. Specifically, his OL treatment generated a linear zigzag outcome curve as a function of angle difference, while the NL model gener- ated a sinusoidal curve. The difference opened the possibility of experimental test. Later treatments extended this approach to populations of photons pairs correlated in polarization but with similar ex- pectations15,20,21, and subsequent experiments, mostly with photons, demonstrating ‘violations of Bell-type inequalities’ have provided the main underpinning for the non-local picture4,15,19,21,22. The resulting world-picture, with properties of entangled quantum states shared and dispersed though space-time, has excited physicists, puzzled philosophers, inspired sci-fi lovers, and emboldened mys- tics over the last half-century23. Much of the excitement comes from the prospect of new physics, be-

4 cause, as is widely recognized5,20,24-32, current explanations imply an observer-dependent reality gov- erned by laws that permit information-sharing over space-like distance forbidden under classical con- straints33,34. In the first part of this commentary, I discuss this problem, and show that the OL model on which distinction has been based is flawed. I demonstrate an alternative model using populations of correlated vectorial photon pairs (vOL, an application of Einstein’s model), which, contrary to expec- tations of Bell’s theorem 7-9,13,20, gives outcome curves that match those of the NL treatment applied to the same initial state. I show that this behavior is natural and fully compliant with locality con- straints, examine why the Bell-type OL treatments have excluded such outcomes, and conclude that they represent poor choices of model and treatment. This conclusion strongly undermines the claim that violation of Bell-type inequalities excludes local realistic models. In the second part, I address the questions now opened as to whether the experimental results claimed to support the non-local interpretation really do, and whether the orthodox NL approach ade- quately explains the results claimed. I question the justification for the conventional NL treatment. I demonstrate that many experimental outcomes claimed to support non-locality, either do not, or are in contradiction with foundational QM premises, but that the natural properties of the vOL model can, within ambiguities in description of protocols, account for most without any contradiction. I dis- cuss extensions of the NL treatment by co-option of Maxwell’s electromagnetic theory to quantum field approaches. I suggest that these cannot be practically applied at the elemental level because photons are neutral and cannot interact through fields. Alternative treatments that represent photons as carriers of momentum can better represent the actions of photons in quantized processes at that level.

PART A. WHERE BELL WENT WRONG The locality constraints introduced by EPR came from their assignment of properties to dis- crete entities (cf.9,19); if properties are intrinsic, information pertinent to measurement can only be ac- cessed locally, and measurement on one of a pair can have no effect on the distant other. In contrast, NL treatments require evolution of entities in a superposition of all possible states, vectorial proper- ties are neither intrinsic nor discrete, and cannot be known until measurement, and, for entangled partners, measurement of one simultaneously actualizes that and the distant other in appropriate cor- relation. Bell11,33 claimed that, under such local constraints, when vectorial correlations were tested as a function of analyzer settings, OL expectations from a population could never match NL expecta- tions13,35.

5 Bell’s original treatment13 considered electrons with opposite spin9, determined on discrimi- nation using Stern-Gerlach magnets. The treatment was refined and extended to photons and polariz- ers by Clauser, Horne, Shimony and Holt15,21 (CHSH), and later adopted by Bell, leading to the Bell/CHSH (BCHSH19,20) consensus that inspired the early experiments22,36-39. An atomic beam (Ca or Hg) was excited by arc discharge22,39, or, in later work, by lasers36-38,40, and pairs of photons with correlated orientation were generated in a two-step decay process (cascade). The pairs evolved to separate measurement stations for analysis using polarizers set to differentiate between expectations. In more recent tests, lasers have been used to excite parametric down-conversion (PDC) in non-linear crystals17,18,41-43, using setups like that in Fig. 1. These latter sources have also been exploited in more complex arrangements following similar protocols, to explore applications in quantum encryption, teleportation and computation, and to test the finer points (loopholes) arising from the entanglement debate16,17,19,44-47.

1. The BCHSH consensus. Before discussing where Bell went wrong, it is necessary to understand his logic. Even those familiar with this field might benefit from a re-cap, so a summary is provided in the SI (Section 1). I also discuss, mainly in the second part, how Bell’s thinking was also determined by his earlier analy- sis of hidden variables. Here, a brief synopsis of the technical aspects will serve to frame the argu- ment. Although Bell initially considered electrons, photon pairs have provided the context for most experimental reports and can be discussed through essentially the same framework. The early treat- ments and experiments 15,19-21 considered photons with states designated by H (horizontal) or V (ver- tical) in pairs correlated in Bell-states HV/VH or HH/VV, and measured coincidence in detection at two stations, tested at four settings of the polarizers in a setup41 such as that in Fig. 1. With α or α' and β or β' as polarizer settings respectively at stations 1 and 2, the approach considers possible coin- cidences at settings α, β; α', β; α, β'; and α', β', chosen to discriminate between outcomes expected, - the OL zigzag and the NL sinusoid. The expectations from coincidences in pairwise measurements at 19 separate stations, (Eα,β, etc.) , then give outcomes summed through:

SBCHSH = Eα,β + Eα,β' + Eα',β − Eα',β' (eq. 1a), where E values reflect normalized coincidence yields at each of the four settings. At the elemental level, a value of +1 was assigned to a photon detected in the ordinary ray, and of -1 if in the extraor- dinary ray, and each of the four Eα,β, etc. terms can have a value ±1. The sum SBCHSH is then con-

6 strained to the range ±2 so that SBCHSH ≤ 2 (eq. 1b). This OL outcome was compared with the expec- tations from the NL treatment. Bell’s NL approach is discussed in greater detail later (Part B, section 5), but started from vector projections resulting from operation of the Pauli matrices on the discriminator vectors: ⃗ ⃗ (휎⃗ ∙ 훼⃗ 휎⃗ ∙ 훽)=−훼⃗ ∙ 훽 = −cosσ (eq. 2). Using photon pairs and standard QM principles, the matrix operation generates a sinusoidal curve as a function of angle difference, σ, between polarizers at different stations, with an amplitude of ±2, invariant on rotation of the photon and polarizer frames. The invariance depends on the dichotomic correlation in spin state in the population, actualization from the superposition as photons with real vectors in alignment with the polarizer frame. Then, the yields at the polarizers are given by Malus’ law, the symmetry of the matrix operations leads to cancellation of contributions from the photon vectors, and coincidences are dependent only on angle difference, σ. Fig. 2 (see legend for explana- tion) shows a unit circle representation of such projections, leading to the observable Malus’ law val- ues at particular settings of the variable polarizer.

Equations (eqs. 1a, b) are taken as the basis for Bell’s OL treatment, and values for SBCHSH then depend on estimation of values for the Eα,β, etc. terms through a conventional application of probability theory. In the seminal treatment, as adapted for photons21, Bell first asked what would be expected from local measurement on a stochastic population. In the context of photons, this gives an elemental probability 푝(휆, 훼), derived from the vectors of the photon polarization, λ, and of the po- larizer, α. The outcome with a population, is then given by integration, 풑(휆, 훼) = ∫ 휌(휆)푝(휆, 훼)푑휆 (eq. 3a), where the function 휌(휆) is a normalized probability distribution for λ. With the stochastic population considered, terms in λ would cancel out, because∫ 휌(휆)푑휆 =1 (eq. 3b) (the isotropic con- dition). As a consequence, no information about λ would be available from measurement on the pop- ulation at either station. OL expectation of coincidences could be obtained by similar integration us- ing the product of elemental probabilities at the two stations. With polarizers set at α (station 1) or β (station 2):

퐸,() = ∫ 휌(휆)푝(휆, 훼)(휆, 훽)푑휆 = ∫ 휌(휆)푝(휆, 훼, 훽)푑휆 = 풑,(훼, 훽) (eq. 4), with similar equations for other polarizer settings. In equating the first and the second equations on the right of eq. 4 (RHE1 and RHE2), and in deriving 풑,(훼, 훽), Bell concluded that, as implicit in RHE2 and the i sotropic condition, eq. 3b would come into play, so that terms in λ would also cancel here. Then, 풑,(훼, 훽), and Eα,β (OL), etc.,

7 and hence SBCHSH, would depend only on polarizer orientations. From this he concluded that photon vectors could play no part in determining the outcome, and therefore developed an OL model in which, instead of the polarization vector, the photon correlations were represented by the scalar sign of the spin. This has been likened to an “exploding penny” model, in which the explosion separates the head side from the tail side, with the two parts hurtling apart in opposite directions48. The equa- tion he suggested,

Eα,β(OL) = p1,2(α, β) = -1 + 2σ/π (eq. 5), yielded a linear dependence on σ in the range ±1, giving a zigzag curve as a function of angle differ- ence, invariant on rotation of frames, yielding an amplitude in the range ±2 for SOL, the sum of coin- cidences at the four settings expected from eq. 1a. Since curves from the two models are different, they predict different outcomes from coinci- dence measurements at canonical angle differences along the curves, giving “inequalities” expressed as limits of either SNL ≤ 2.83 from the NL sinusoid, or SOL ≤ 2 from the OL zigzag. The same limiting value can alternatively be derived generically for any OL model12,15,19-21,49 from substitution of elemental counts of ±1 in eq. 1a above; for the four terms, this gives SBCHSH as ±2 (eq. 1b). Since the mean from integration over a population cannot exceed the maximal elemental value, the limit SOL ≤ 2 for all OL models follows from this approach. These are the first and second of Bell-type inequalities. Despite spectacular technical and theoretical advances, this treatment con- tinues to provide the consensus framework for discussion of the inequalities in the entanglement de- bate. In these terms, a “violation of Bell’s inequalities” means that the result of an entanglement test significantly exceeds the ≤2 limit expected from the OL model. Bell’s third inequality came from an OL model based on a stochastic population of correlated photon pairs retaining vectorial properties and will be discussed in detail later (Part B 1).

2. The critical experimental results. Over the years following the first results22,39, an early consensus developed based on out- comes testing readily distinguished features; the difference between Bell’s zigzag and the NL sinus- oidal curves, and the invariance of the sinusoidal outcome on rotation of the polarizer frame4. With cascade sources, technical difficulties and inefficiency of detectors (at <20% efficiency, <4% of co- incidences could be detected), required normalization of coincidence counts, and correction for “ac- cidental” counts22. The ‘inequalities’ were represented in the δ-function, which scored the difference between coincidences measured at canonical angle differences, and those expected from the OL

8 limit22,36,38. In terms of the above criteria, the NL case wins hands-down, because the OL model was shown to fail; with one exception, later disavowed (cf.50), early tests22,36,37,39,40 showed the rotational invariance expected from either model, but the sinusoidal curves expected from NL, rather than the OL zigzag. After normalization and correction, amplitudes were interpreted as showing the full visi- bility expected from NL models4,20-22,28,36,48, and these results steered the debate towards a consensus supporting Bell’s theorem49. Subsequent experiments of increasing sophistication16-18,41,51-53 using PDC sources, have refined results so as to approach closer and closer to the NL expectation of 2.83, with the ‘violation of Bell-type inequalities’ scored through standard deviations from the OL expec- tation of ≤2 (cf.36,38,41,47,54). In recent work with high-efficiency detectors (~91%), outcomes meas- ured have exceed the OL expectation without any special corrections55,56. Based on such results, the quantum theorists have concluded that no explanation in local realistic terms could be tenable. For Bell13, “…the quantum mechanical expectation value cannot be represented, either accurately or arbi- trarily closely…” by any equation in the form of eq. 4. Since in all reports the results matched the NL sinusoid, they could only be accounted for in local realistic terms by invoking “hidden variables”, re- quiring “conspiracies” between source and stations to explain the vectorial outcome, a view summa- rized in “…if [a hidden variable theory] is local it will not agree with quantum mechanics, and if it agrees with quantum mechanics it will not be local…” (from Bell57, as qualified by Shalm et al.56). Various scenarios (“loopholes”19), most notably detection and communication loopholes, are dis- cussed through which such conspiracies could be enabled. Experiments, some of spectacular sophis- tication (cf.55,56,58-60), have successfully eliminated such loopholes, apparently solidifying the NL case. Shimony’s conclusion19 that “the prima facie nonlocality of Quantum Mechanics will remain a permanent part of our physical world view, in spite of its apparent tension with Relativistic locality” is widely accepted (cf. 39).

3. Simulation of a vectorial OL model (vOL) shows that the BCHSH OL model is inade- quate The BCHSH algebra seems impeccable, but something is not right. Prediction of the sinusoi- dal outcome of the NL treatment depends on application of Malus’ law to pairs actualized with real vectors. In the derivation of Bell’s OL zigzag above, vectorial information about the photon source had been excluded, so Malus’ law was not involved. His model could never generate sinusoidal curves because the vectors needed at the polarizers had been replaced by scalar spin states.

9 In order to better understand what factors might be important, typical tests17,41,53 using photon pairs (Fig. 1) have been modeled through discrete math simulation. The initial intent was to address the naïve question implicit in the above paragraph: What would be the outcome if a vectorial local realistic (LR) population of correlated photon pairs was analyzed using Malus’ law compliant polar- izers? The short answer is – the sinusoidal outcome expected from NL predicates (Figs. 2, 3A).

(i) Simulation program. The core of the program is simple. For each ‘experiment’ a population of ‘photon’ pairs is generated with each partner defined by an explicit polarization vector appropriate to program settings for source, Bell-state, correlation, etc. (the Make Light subroutine). Uncertainties are implemented where appropriate (value of photon angle for each pair in a stochastic population, variable polarizer settings, allocation of photons to beams, etc.) by use of a random number generator. Orientations of the polarizers are set after the photon population has been established (“with photons in flight”), but before discrimination, measurement and plotting of coincidences and singles counts (the Measure- ment and Plot Point subroutines). The most important difference from Bell’s scalar dichotomic model is that the photons of the vOL population measured have the intrinsic vectors Einstein would have expected. As in real experiments, yields at the ‘polarization analyzers’ are derived from discrete meas- urement of elemental outcomes; a photon appears in either the ordinary or extraordinary ray of a po- 2 larization analyzer. Discrimination at the polarizers is assumed to follow Malus’ law (I/I0 = cos φ, with I/I0, the normalized transmission, θ the orientation of the polarizer, λ the polarization vector of the photon, and φ = θ - λ) At the elemental level, Malus’ law applies statistically; the yield calculated is a probability for transmission corresponding to the BCHSH term, 푝(휆, 훼) above. The sampling of the statistical spread is implemented by incrementing counts at detectors (Q or R at station 1, S or T at station 2, Fig. 1), based on comparison of the normalized Malus’ law yield to a random number between 0 and 1; if the yield is greater, the photon goes to the ordinary, else to the extraordinary ray. This gives the single counts. Two exceptions to this rule are in simulation of Bell’s OL model, and in mimicking the QM expectations. Simulation of Bell’s OL model is by comparing the calculated yield to 0.5 to give a binary outcome (see below); simulation of the QM expectations is by implementing the propensities of each photon implicit in the outcome of the matrix operations as summarized by Shimony19 (see Part B, Section 7). For each pair, coincidences between stations are scored, with the mean plotted to give points on the coincidence curve. Scoring of coincidences can be selected from a

10 choice among four different algorithms. With the two exceptions noted, all these components are fully compliant with strict locality constraints. At any particular setting of controls, a point on the outcome curve is determined from a popu- lation of pairs by a simple count of pairwise coincidences. Useful features of the program are diag- nostic aids. The menu bar (Fig. 2, top) includes different Run options that implement a complete set of experiments generating an outcome curve. These Run options iterate though three subroutines (Make Light, Measurement, Plot Points in the Menu bar) that implement the functions suggested, to generate a point on the curve. The user can choose from several different counting algorithms by as- signing values (count increment, true/false, ±1 in CHSH protocol) appropriate to the different algo- rithms (anti-correlation count, Boolean coincidence count, CHSH count. An emulation of the Freed- man and Clause (FC) δ-count, can also be chosen (see Program Notes for implementation). Which- ever is used, the outcome is, as appropriate to the settings, essentially independent of the algorithm. Counts at the elemental level follow the complementary symmetry expected12,19, seen experi- mentally37,38,40,61, and discussed at length by Mermin61. Elemental counts for the last population tested, and vectors relevant to measurement of each pair, can be displayed pair-by-pair in the Gadgets (top right, Fig. 3A). The pairwise counts generated at a particular setting of the polarizers show ele- mental values, - either +2 or -2 for the CHSH count, 0 or 4 for the anti-correlation or Boolean counts, which accumulate when integrated in a population to give points on a curve plotted as a function of polarizer angle difference, with the Boolean count 90o out of phase with the anti-correlation count. The range shown in the Gadgets (±2 or 0-4) might lead to some confusion, because, on a sin- gle run, the fixed polarizer is set at a particular value, and the range of the count could reflect only ' two (for example α, β and α, β ) of the four E terms contributing to SBCHSH (a range ±1 or 0–2). The range shown in the Gadgets comes from scoring both detection and non-detection, which is practica- ble with perfect detectors, and facilitates taking cross-products, but counts each coincidence twice (half of the coincidences are redundant because a non-detection always mirrors a detection). How- ever, after normalization to the two photons of a pair, the range at a particular setting (±1 for the CHSH count, etc.) is the same as the conventional scoring, and this is the default for outcome curves shown in the display panels. In each of the counting algorithms, the summation over a population represents an integration in the form of eq. 4, in which RHE1 is considered as the expectation. The outcome is determined solely from the elemental counts. The program functions at the level of observables, defined by the use of Malus’ law probabilities. No algebraic sophistication is involved in the simulation, - refractive components are treated naïvely in terms of their empirical behavior, and only is

11 considered. Elementary trigonometry is used, but in only two contexts; (i) to determine from Malus’ law what yields would be expected at the settings used, so as to implement the statistics above; and (ii) to calculate Malus’ law compliant theoretical curves. Links to the executable program (Bell_Ineq.exe), source codes (in Microsoft Visual Basic 12), and Program Notes (also accessible as Help in the program), are available in the SI.

(ii) Program outcome. The correlations observed depend on the model for the photon population, on how the ‘polarizers’ are chosen to respond (Fig. 3 A-E), and on choice of settings. When the polariz- ers are set to implement Malus’ law, photon pairs are oriented in a common frame, and share an an- gle with the fixed polarizer, coincidence counts from a vOL population follow the full-amplitude si- nusoidal curve expected from the NL treatment (Fig. 3A, LR0 curve), not the linear zigzag expected from Bell’s OL model. Compared to the coincidence count, the CHSH count gives essentially the same curve (Fig. 3C) but offset because of the choice of ±1 for elemental values, and the anti-correla- tion count is 90o out of phase. The curves show the full amplitude (‘visibility’) expected from the NL treatment. This LR0 outcome applies at all angles for the common frame (Fig. 3 A, C, E), - a classi- cal rotational invariance. When the photon pair orientation is random (isotropic in the plane of measurement), and the polarizer function is natural, the outcome is a sinusoidal curve, which shows the invariant behavior expected by NL on rotation between the photon and polarizer frames, but with amplitude half that expected from NL (Fig. 3B). Such behavior (LR2 curve) is fully compliant with locality con- straints62,63; the reduced amplitude represents the entropic penalty on measurement of a stochastic source. As discussed at greater length in Part B 1, it corresponds to expectations from Bell’s third in- equality. With the photon source randomized and the polarizers set to Bell binary mode (equivalent in effect to Bell’s scalar choice), the correlations follow the linear zigzag expected from his OL model20, with a partitioning proportional to σ (Fig. 3D, LR1 curve). This outcome is also invariant to rotation of frames. However, in contrast to the LR2 outcome, it shows full amplitude. Note that the photon population of the simulation differs from Bell’s original model in retaining its vectorial char- acter, so the outcome here depends on forcing the polarizers to behave (unnaturally) as binary rather than Malus’ law discriminators (see above for implementation). Outcomes from other combinations are unremarkable, though not without interest. The user can set parameters for modifications to the photon state on insertion of additional refractory elements into one or the other path, or both. Different coincidence counts give the result appropriate to the

12 phase relationships tested, with different Bell states showing the expected behavior. However, with oriented photon pairs, when the polarizer frame is rotated away from alignment with the photon frame (or vice versa), the amplitude of the sinusoidal curve, as expected, decreases, and is zero when the rotation is by 45o (Figs. 2, 3E). This is in contrast with NL expectations, where full-amplitude is expected at any orientation of the photon frame, including stochastic. The NL outcome can be gener- ated by implementing a “QM simulation option” to match the effective outcome suggested by Shi- mony19, - as discussed later in Part B, this is in effect, a Maxwellian demon. For settings at which Bell-correlated pairs were analyzed, the local mean yields (the ‘singles- counts’) at the four outcome rays have the same property, - a mean probability of 0.5 (within the sta- tistical limitations of the sample), independent of photon source or polarizer settings (right-panels in Figs 3A, B, D). This is what Bell expected. An Analog option allows display in the right panel of the fractional yield differences given by Malus’ law (the sinusoidal curves in Fig. 3C, right panel, explained in the legend). The theoretical curves in the left panel can be derived from the analog distributions in several ways, but the points plotted always reflect the statistical outcome from counts of pairwise coincidences at the two sta- tions. The full visibility theoretical curves expected from the NL treatment (in effect the LR0 curves) are displayed in the left panel when frames are aligned to match that expected under NL predicates. In summary, the outcome shown in Figs. 3A and 3C are the full-amplitude sinusoidal curves expected under NL predicates (cf.16,18,28,48), generally presented as impossible from any OL treatment, but which here demonstrate a violation of Bell’s inequalities from a vectorially correlated population of photon pairs, fully consistent with OL constraints.

4. What can we learn from the simulation? Local realistic models are often presented as classically constrained and are contrasted with NL models following QM predicates. However, in the argument between Einstein and Bohr both models were QM inspired, though interpreted from different perspectives. The behaviors extracted in my vOL simulation follow from empirical treatments that are essentially classical but are quantized as expected by Einstein. No model following his expectations would be antithetical to foundational QM principles. In the simulation, all interrogative interactions (refractive discrimination at polariz- ers, half-wave plates (HWPs) etc.) involve discrete photons and full commitment of their action. In effect, for entities with intrinsic properties, all exchanges are quantized and local, but all behaviors (except the QM simulation and Bell binary options) follow empirically justified laws (see Part B for discussion).

13 The vOL approach depends explicitly on the factorizability of cross-probabilities within a scalar algebra64 (eq. 4, RHE1), generally considered as characteristic of valid implementations of the locality constraints19. The LR0 outcome shows, contrary to conventional expectations, the same full amplitude sinusoidal curve as the orthodox NL treatment. Such an outcome should not come as a surprise. For any particular setting of the fixed polarizer, there must always be a vectorial LR popu- lation (readily available using PDC) that generates the same curve as NL, - that in which the photon population is oriented and aligned with the polarizer frame, - the configuration expected under NL predicates on actualization at the time of measurement. The outcomes are then the same because the vector projections are the same (see Fig. 2). As long as the photon vectors are represented, the only condition needed for prediction of such an outcome is the alignment of frames. An important conclu- sion from this is that the difference between treatments must lie in how alignment is achieved; it is implicit in the NL matrix operations but has to be explicit under vOL constraints. With this proviso, since the curves are the same, the differences in yields at canonical values (~2.83) ‘violate Bell-type inequalities’ just as does the NL value. The conclusion that no OL outcome can match the NL expec- tation value must then be invalid; in particular, the OL limit of ≤2 must be artificial. Since the sum- mation of outcomes giving the LR0 curves is equivalent to integration in the form of eq. 4 (the RHE1), Bell’s conclusion that “…the quantum mechanical expectations cannot be represented (in that form) …” is also clearly wrong. I make no claim for originality in introducing the vOL model; it is simply Einstein’s perspec- tive applied in a vectorial context. Several previous efforts (cf.62,65-72) have arrived at similar conclu- sions, most explicitly in62,67,73. The earliest of these by Angelidis65, a protégé of Popper74, was dis- missed by Garg and Leggett49, essentially in terms of the limit of ≤2 from the second (BCHSH) ine- quality (eq. 1 above). Their brief paper was selected by the editors as representative of a much wider community that responded similarly. This consensus reflected a confidence in the expectations of Bell’s theorem, reinforced by its apparent validation in contemporary experiments22,36-38,40. The rejec- tion was perhaps understandable in a historical context, reflecting wide acceptance of the Copenha- gen interpretation. Similar dismissals of all later claims have been justified by the same rationale. However, from the above, perhaps confidence in this dismissal was misplaced.

5. Bell’s three mistakes

Note that in Bell’s analysis summarized above, the value of SBCHSH depends only on the angle difference between polarizers, σ. The properties of the photons in the source population were not

14 considered because he concluded that the vectors were “hidden variables”35, and could not be ac- cessed locally. Without vectors for the photon source, discussion is constrained to Bell’s framework, - a mental box that has effectively precluded consideration of Einstein’s model. a) Local measurements. For stochastic sources, as Bell pointed out, the mean local yields would necessarily be independent of polarizer orientation. All early treatments involved such 22 sources; the outcome was measured in the R1/R0, etc., terms of Freedman and Clauser , and this was use in later reports using Ca-cascade sources, and also by Fry and Thompson36 with a 200Hg-cascade excited by laser. It is simulated in the singles-counts of 0.5 (see right panels of Figs. 2A, B, D), and is diagnostic of an isotropic condition in the plane of measurement. Bell’s first mistake was to draw the wrong conclusion from this outcome, - that vectorial properties of photons could not be involved in the behavior seen on local measurement, and should therefore be excluded consideration. b) The singles-count at each station are accounted for by natural behavior at the quantum level. Although values for individual photon vectors are lost in the mean, the behavior observed locally must access them at the elemental level. It depends on what happens at the polarizers, where the Ma- lus’ behavior law requires vectors for both photon and polarizer. The experimental outcome can then be explained naturally in terms of local vectorial properties. For elemental measurements, I/I0 (the

Malus’ law expectation, see Section 3, (i)) gives the BCHSH probability, 푝(휆, 훼), and on integration over a population at λ, the Malus’ law yield. With a stochastic population, and a polarizer at any set- ting θ, sampling λ by integration over the hemisphere (as in eqs. 3a, b) would show a distribution of values for 푐표푠휑 varying with λ, centered at the polarizer vector, θ, and with the mean yield of 0.5. Since the same curve is found at all values of θ, this accounts in terms of local properties for the be- havior Bell took as demanding exclusion of such properties. There are no “hidden variables” in this treatment, so their invocation in discussion is not useful, and Bell’s conclusion that vectorial proper- ties were not involved in the behavior observed was wrong. The zigzag outcome is contrary to the sinusoidal Malus’ law behavior demonstrated in 200 years of experimental work exploring polariza- tion. When using an oriented source of photon pairs with dichotomic distribution of H and V in each population (as in Bell states HV/VH or HH/VV from PDC), the mean yield in any ray is given by 0.5(cos2(θ – λ) + cos2(θ – (λ + 90o))) = 0.5(cos2θ + sin2θ) = 0.5 (eq. 3c), where θ is the orientation of the polarizer and λ that of the photon reference frame (the H photon)57,69. The partition of H and V

15 photons is isotropic because symmetrical about the reference axis, and gives in the mean, the same yield as Bell’s integration for a stochastic source. Within the above constraints the outcome is there- fore independent of source model, or specific values for θ and λ. A similar mix could be expected from either an OL or a NL state, so this outcome is of little interest in distinguishing between models. c) Expectations from comparison between stations. No vectorial correlations between stations could be predicted from the mean yields from local measurements because all information that would allow comparison of each photon to its partner is lost in the mean. Experimentally, correlations are determined from pairwise comparison of elemental measurement outcomes at separate stations. This selection is important because with pairs in stochastic orientation, correlations are conserved only on a pairwise basis. With cascade sources, the protocol was designed to select pairs by temporal coinci- dence, and choice of color filters based on properties expected from the cascade, and on direction of flight. Coincidences were maximal when polarizers were aligned, demonstrating that both photons of a pair had close to the same orientation; since these last two properties are expected from conser- vation of angular momentum in the source process, the protocol was predicated on determinate prop- erties. Correlations were either detected on-the-fly by coincidence counters, or determined from data recorded and time-tagged on-the-fly and analyzed later. Further analysis requires the pairwise data, their time of measurement, knowledge of polarizer settings, etc., but all this information is exchanged subluminally 68,69. Different approaches providing justification for the OL limit of ≤2 were outlined above: (i) Bell’s derivation from the zigzag. In deriving the zigzag13, Bell started by applying eq. 4 a stochastic source. Then, in light of RHE2, the integration through eq. 3b eliminated con- sideration of vectorial properties. These were replaced by the sign of the spin, leading to a scalar partitioning to different hemispheres (eq. 5), equivalent in effect to the binary discrimination giv- ing my LR1 curve. Peres48 suggests that “Bell’s theorem is not a property of quantum theory. It applies to any physical system with dichotomic variables, whose values are arbitrarily called 1 and -1”. However, EPR certainly considered their model to be consistent with the foundational QM principles established by Einstein over the previous three decades. If Bell’s OL treatment was “…not a property of quantum theory…”, it was because, contrary to EPR, he stripped the entities of their intrinsic vectorial properties. He thereby eliminated consideration of Einstein’s model, - Bell’s second mistake. The zigzag (as demonstrated in the LR1 curve in my simulation, Fig. 3D) would be expected for radiating scalar partners from dichotomic pairs (the “exploding penny” model), but scalar correlations could never lead to a sinusoidal curve.

16 (ii) Expectations from eq. 4 depend on the degree of order in the source. With a stochas- tic source, the outcome expected from eq. 4 is dependent only on the polarizer difference; it is independent of the setting of the reference polarizer. This behavior is a consequence of the iso- tropic condition, which justifies invocation of eq. 3b., and leads to an expectation that the out- come, 풑,(훼, 훽), would be rotationally invariant. However, this is not a consequence of any loss of vectorial properties. This can be seen in the fact that all experimental reports have showed si- nusoidal outcome curves; these simply shows Malus’ law behavior at the polarizers, which re- quires that photons have vectors. This natural behavior is also reflected in the half-amplitude LR2 curves of the simulation, and their rotational invariance. The apparent “loss” of vectorial conse- quence is a measurement problem, a trivial epistemological issue. In this light, Bell’s third mis- take lay in extending the wrong conclusion from his first mistake (above) to interpretation the outcome expected from the count of coincidences. Both for the singles counts and the coinci- dence counts from an isotropic population (eq, 4), he interpreted the expectation that the behavior would be rotationally invariant as showing that the vectorial properties of the photons could not contribute to processes determining that behavior. His replacement of the vectorial property by the sign of the spin was an ontic surgery, in effect removing the vectorial property. This made it impossible to take into account the elemental vectorial processes involved in the behaviors. The count from integration at a single station and the integration of the pairwise differ- ences between stations involve different operations. The singles-counts come from integration at one station of elemental responses from an isotropic mix of V and H photons (eq. 1-3). Then each population measured at separate detectors gives the same singles count, no matter how the polarizer is set (Section 5 b)). In contrast, pairwise measurement involves detectors at two sepa- rate stations, and analysis in which each photon is compared (via Malus law) to its space-like separated partner. While in the singles counts, summing the elemental probabilities at a local sta- tion (eqs. 3a and 3b, or 3c) gives in the mean the same 0.5 local yields at any polarizer setting, when applied in pairwise counts, the same elemental probabilities lead to Malus’ law differences which accumulate to yield, in the mean, points on a sinusoidal curve. The Malus’ law outcome is 2 2 given by Eα,β = cos σ - sin σ = cos2σ, etc.; at the elemental level, the terms are probabilities ex- pressed in the distribution of ±1 values on detection at each station. In the mean from a popula- tion in a stochastic distribution of vectors, this generates a curve of half-amplitude, analyzed in detail in Part B.

17 With an ordered population, since eq. 3b does not come into play, although RHE1 still holds, the RHE2 term is no longer relevant, and expectation of rotational invariance from

풑,(훼, 훽) no longer holds. With an ordered and aligned photon source, the mean from integra- tion of pairwise coincidence (in effect using RHE1) gives the LR0 curve, but the same outcome can also be derived analytically from that equation, using the four cross-products between yields calculated at the two stations (QS, RS, RT, and QT) taken for each configuration of the pair (for example VH or HV). The outcome (the green curves) depends on the degree of order in the popu- lation. For ordered populations, with the photon and polarizer frames aligned, projections from the eight comparisons above lead to the same Malus’ law outcomes as in the matrix operations of the NL approach, and differences give points following the full-visibility curve expected from this (see Fig. 2 and legend). For a stochastic population, the value for λ for each particular pair will be different. In pairwise measurements, partner is still compared to partner, but the mean will be reduced by the entropic penalty from cancellations arising from the stochastic distri- bution of values for λ to give a cos2σ curve of half-amplitude (the LR2 curves are analyzed in more detail in Part B 9). With ordered populations misaligned, the standard probability ap- proach above gives the green curves in Figs. 3E, the Malus’ law result. In all cases, the sinusoidal shape of the curve simply shows Malus’ law in action, nothing more. No natural vectorial state could generate the zigzag; no scalar state could generate the si- nusoid. (iii) Derivation of the <2 limit from the elemental count, - the BCHSH inequality. This approach, first hinted at by Clauser et al.15, further developed by Bell20,35,75 and by Clauser and Horne21, expressed with clarity by Leggett12,49 and reviewed comprehensively by Shimony19, was based on showing from the elemental coincidence values of ±1, that Eα,β + Eα,β' + Eα',β − Eα',β' from eq. 1 is limited to ±2, and noting that this constrains the mean value from a population, SOL, to ≤ 2, to set a limit for all OL models. This was then compared to the SNL = 2.83 value from the NL treatment at canonical angle differences. The maximal value of 2 defines the OL limit, the NL expectation of ≤2.83 unambiguously exceeds that limit (both features are also demonstrated in the simulation), so the conclusion that no OL model that conforms to Bell’s constraints could match the NL expectations might seem unassailable12,19. However, although the math is correct, the conclusion is wrong. It depends on two mistaken assumptions: (1) that the maximal value of

2 applies only to the OL case, and (2) that it is directly comparable to the canonical value for SNL of 2.83. That neither is true becomes obvious from examination of the outcome curves scaled to

18 four units through the S parameter (Fig. 4). The same sinusoidal curve in the range ±2cos2σ can be generated from either the vOL or NL model, (disproving (1)); and, although both the maximal value of 2 and the NL limit of 2.83 belong to the same outcome curve, they describe different properties of the curve, so are not comparable (disproving (2)). These two mistakes cannot be blamed on Bell; they have been perpetrated through their acceptance by the whole commu- nity12,29,48,49. (a) The range ±2 is a consequence of the choice of elemental values of ±1. Since cosσ varies between ±1, the same elemental values and the same maxima and minima also define the NL curve; as shown in the simulation, at appropriate alignment, the NL and vOL curves are the same. The maximal value and the properties of the curve therefore apply to both models. The S parameter can take any value in the range ±2 (from eqs. 1a, 1b):

-2 ≤ S = Eα,β + Eα,β' + Eα',β − Eα',β' ≤ 2 Although the count constrains the maximal amplitude of the outcome curve to 2, the value of a point is constrained by the ±2 limits; any difference between two points is constrained to the 4 unit range of the curve. (b) That the OL limit of ≤2 is problematic should then be obvious. The OL limit of 2.0 comes from a singular point, the maximum of the curve. In contrast, the NL limit comes from differences between two points on the curve. At any pair of canonical values for σ, those two points fall symmetrically about 0 within the scale range of ±2 (right scale pertaining to the CHSH count in Fig. 4). For example, at canonical settings of 22.5 and 67.5 (as in Fig. 4), the S values are 1.414 and -1.414, with the difference of ~2.83 applying to both models. Exclusion based on the comparison between the maximal value and the difference is then obviously absurd. (The points on Bell’s zigzag are 1.0 and -1.0, giving a limit of 2 from the first inequality, but since the model is wrong, this is now of historical interest only.) (c) For any simple count of coincidences (Fig. 4, left scale, the anti-correlation count (red points) or the Boolean coincidence count of Fig. 3A), the curve will fall naturally in the range 0 - 4, and any ordered population in which frames align will give the full-visibility sinusoidal 4cos2σ curve. Values at the canonical intercepts show the same difference ≤2.83, but this is un- remarkable when compared to the amplitude limit of 4.

Limits of <2 as derived above provide no basis for discrimination between local and non- local models, - they reflect instead either a poor choice of OL model or a poor treatment or both. When the photon and polarizer frames are aligned, there is no difference between vOL and NL

19 expectations. The failure to find Bell’s OL expectations experimentally is unremarkable since it was based on an unrealistic model. d) Fitting the sinusoidal curves. When real vectors in distinct frames for photons and polarizers are used to represent values pertinent to a vOL model, the results differ from NL expectations only when the frames are misaligned. For aligned populations, the full-visibility sinusoid can be derived directly from the Malus’ law yield differences in each ray (right panel of Fig. 3C and legend). The conventional NL treatment (cf.19) in effect actualizes photons with their vectors in that same alignment. When oriented OL populations are not aligned, the more complete probabilistic analysis giv- ing the green curves (Fig. 3E) is required. In the simulation, such curves are obtained (as a function of σ) by counting the pairwise coincidences, finding the mean by the integration of eq. 4, using

RHE1, and calculation of SvOL from the sum of eight outcomes as discussed above (Part A, 5a)). An- alytically, the green curves are derived from the same eight outcomes but calculated from the Malus’ law expectations. These treatments embody all the BCHSH realistic prescriptions for OL treatments and generate simulated or theoretical curves that fit all outcomes using oriented populations. If detec- tors in both outcomes of a polarization analyzer are measured, the same eight terms contribute to the mean count from experimental pairwise comparisons. The set is formally equivalent to those in play from the matrix operation of the NL treatment applied in the plane of measurement (Fig. 2). e) Known unknowns. No elemental measurement can lead to complete specification. Elemental events involving photons are necessarily quantized, but classical statistics will still apply (see76, and Part B). The simulation demonstrates that the sinusoidal outcome curves are not a consequence of indeterminacy, or of inseparability, superposition, or any of the algebraic paraphernalia said to be re- quired to deal with QM uncertainty. As long as the “uncertainties” are distributed normally about the mean, the counts of elemental pairwise coincidences will be sufficient. Attribution to the ‘entangled state’ of ‘super-correlations’48 on the basis of the sinusoidal curve is nonsensical; the shape of the curve requires nothing more than Malus’ law operating on pairs in vectorial correlation (discrete in the vOL model, LR bivectors in Clifford algebra treatments70,77-79, or the pairs actualized with aligned vectors in the NL treatment).

The vOL model involves no “hidden variables”. All the information needed to account for the outcome is carried as intrinsic properties of discrete quantized entities; the information arrives with the photon. There is then no need for “conspiracies” to explain the results; the model is immune to

20 closure of communication ‘loop-holes’ (discussed at greater length later). The only requirement is for the behavior at the discriminators to be natural. On the other hand, the full-visibility amplitude de- pends on alignment; the difference in the two approaches as to how that is achieved is a separate is- sue.

6. Why are these conclusions important? The literature is full of claims (cf.12,15,17,19,20,49) that no objective local theory could yield a curve that departs from the limit of ≤2. Indeed, that limit is nowadays the de facto criterion used for evaluation of the success of experiments in supporting the non-local picture, as shown by claims to “…have measured the S parameter… of Bell’s inequalities to be 2 < S < 2.83, thus violating the clas- sical value of 2 by n standard deviations…” or similar41,47,54-56,80. These comparisons, including in recent “loophole-free test of local realism” using electrons60 or photons55,56,59, are more or less worth- less because the target of 2 (or similar55) is irrelevant. Examples in the popular literature that justify the NL picture based on OL models that use scalar dichotomic qualities such as colored socks81, live and dead cats82, hard/soft, black/white, red/green properties28,61,83 etc., or even the polymorphic quan- tum cakes84, serve only to confuse. Such properties are simply inappropriate to the vectorial states involved and could never generate sinusoidal curves. The simulation highlights a general problem with the BCHSH approach, - that Bell’s OL model (and related treatments) are unable to represent a state with the vectorial properties needed in application of Malus’ law. The inequalities so far dis- cussed show only that no model that omits vectors can generate the sinusoidal curves claimed as sup- porting NL expectations. Justification for non-locality must be otherwise demonstrated.

PART B. MAYBE EINSTEIN GOT IT RIGHT A philosophical hurdle. A newcomer engaging with the entanglement community quickly learns that acceptance of the NL case is general and is based on a conviction that Einstein lost the ar- gument with Bohr; quantum uncertainties preclude assignment of intrinsic properties, and therefore require a treatment in which entities evolve in a superposition of indeterminate states, which are actu- alized only on measurement. The strength of the case was apparent in development of the H-atom model; as atomic spectroscopy provided energy levels for electrons, the potentials for occupancy of orbitals were mapped, completed on inclusion of the spin states in the Schrödinger wave equation, and then more widely applied to flesh-out the periodic table. The standard Copenhagen interpreta- tion, and its ancillary orthonormal treatment became deeply embedded in the zeitgeist of quantum

21 physics. Superposition requires a non-local framework, necessarily treated in the wavy domain. Ad- vances over the 85 years since EPR represent an academic heritage through many generations of a success that has spawned Nobel laureates galore, revolutionized physics and chemistry, and fathered many of the innovations that drive our modern economies. Bell’s treatment was brilliantly framed within this tradition, and the experimental validation of his theorem in the '70s and '80s cemented the non-local view, and extended a confidence that similar treatments should also be applied in the wider spatial context. Insofar as the conventional treatment relates to condensed systems and/or atomic scales, I have no argument with the main conclusions from this spectacular record. However, tests based on entangled photons require their evolution to space-like separation before measurement, and non-local effects are then consequential9. Even Bell expressed misgivings about this side of his theorem81, and confidence must now be further eroded by the wonkiness of the stool on losing two of its legs. The ‘failure’ of local realism arises from the fact that its constraints are real, local, and second law com- pliant. Given the OL model he constructed, Bell’s conclusion seemed justified, and his ideas gained traction because that model was uncritically accepted. The zeitgeist trapped him, and apparently the community in general, in a box that excluded Einstein’s model. Despite advances in physics, the same justifications are still applied, so I will first examine the conventional NL case, and then see if any paradox remaining can be resolved in light of progress beyond the Copenhagen interpretation. In the vOL model, quantum scale entities carry properties intrinsic to discrete states, - Ein- stein’s ‘elements of reality’2,83. Einstein’s case is well known from his criticism of the Copenhagen interpretation as incomplete1,85, but this argument had been discounted by von Neumann who proved that “hidden variables” were not needed in the Copenhagen approach. It was from his analysis of von Neumann’s case that Bell became interested in the entanglement debate (see Section 2 below). Bell himself had earlier14 recognized in the context of the “hidden variables” debate that “…if states with prescribed values...could actually be prepared, quantum mechanics would be observably inadequate....”. Such states are now available in the photon pairs generated from PDC sources. The phase-matching is determined (prescribed) by conservation laws, orientation is deter- mined by the pump laser, and, for example in type II PDC, the two emergent beams are empirically demonstrated to be orthogonally polarized. Paradoxically, the manipulation of these populations in state preparation is explicitly based on full knowledge of their determinate properties. In Sagnac in- terferometer applications58,59,86 (see SI, 3 ii) a) - c)), the polarized components of both PDC outputs (from passage clockwise or counter-clockwise through the interferometer) are separated by beam

22 splitters, and used without mixing to provide distinct polarized populations used in measurement. But use of these is forbidden when an indeterminable superposition is invoked for entangled states. Tak- ing to heart Bell’s message that states with “prescribed values” are contrary to QM, in this section I re-examine the conventional NL treatment and its experimental support, and find it wanting.

1. Bell’s third inequality provides a discriminating case My vOL model is simply an extension of Einstein’s idea. In one sense it is trivial, - if a model matches the alignment expected from NL, of course it will give the same outcome. However, its con- sequences have apparently not previously been appreciated. Otherwise, it would have been obvious that the limits of ≤2 could not exclude local realism. Invocation of that limit would then represent de- liberate obfuscation. Since that is anathematic, whatever has been discussed is something different. In Part A, I show that two of the legs claimed as supporting this stool actually provide neither support nor justification for exclusion of local realism. This failure might be expected to worry the entanglement community. Since the limit of ≤2 does not itself exclude local realism, what does? In discussion with colleagues, loss of the two legs has been dismissed as uninteresting, specifically be- cause of Bell’s third inequality. He originally framed this through “…consider the result of a modi- fied theory…in which the pure singlet state is replaced in the course of time by an isotropic mixture of product states…”, for which he suggested a correlation function, (-⅓a∙β). Although somewhat sketchy, this value is a diminished version the vectorial outcome expected under NL predicates (eq. 2), and involves electron pairs evolving in stochastic orientation, leading to a reduced amplitude compared to the full-amplitude curve 13,14,75. In contrast to the model giving the zigzag, the pairs here retain vectorial properties. A similar reduced amplitude is shown for vOL photon pairs in the LR2 curves generated with a stochastic photon source (Fig. 3B). Leggett (personal communication) has suggested a concise expression of the distinction aris- ing from Bell’s third inequality through the following cases, in which θ1 and θ2 are orientations of the fixed and variable polarizers, respectively:

Case 1: For any possible choice of θ1, there exists an OL model, T, such that for all θ2, fT(θ1, θ2) = fNL(θ1, θ2).

Case 2: There exists an OL model, T, such that for any possible choice of θ1, and for all θ2, fT(θ1, θ2) = fNL(θ1, θ2). In framing these cases, Leggett seems to have recognized that the LR0 outcome of my simu- lation demonstrates model T for Case 1; I take this as a partial validation of the conclusions in the first part of the paper. Failure of the vOL treatment to predict model T for Case 2, the full-amplitude

23 rotational invariance, leaves that as the remaining justification for exclusion of local realism. Despite claims from others to the contrary71,77 (see SI, section 3 (iii)), my simulation shows, in agreement with Bell, that constraints from local realism mean that no OL model can predict the NL rotational invariance19,87. Leggett’s distinction omits an important consideration that will figure in further discussion, - the vectorial properties of the photon source apparent from the generating transition. Though ex- cluded, and therefore of no relevance in the NL case, these properties have to be considered in any vectorial realistic model, and any comparison has to include a dissection of their fate in the NL case. For photons carrying intrinsic properties, Case 1 then becomes more highly restricted, limited to the situation in which the photon population is ordered, and its reference frame is aligned with θ1. For Case 2, the outcome predicted remains unconstrained, in the sense that the NL outcome is independ- ent of the initial orientation of frames, or of whether the population is stochastic or ordered. Any pop- ulation of pairs with propensities in dichotomic correlation must give the same full-amplitude rota- tional invariance for any reference frame at the polarizers (see below). However, since in the stochas- tic case, the initial state is clearly disordered, and the outcome observed depends on actualization of an ordered and aligned photon state at the polarizers, to be credible the treatment would have to in- clude a mechanism, including a work-term, to account for the ordering. I argued below that this re- quirement cannot be naturally satisfied.

2. How credible is the NL case for non-locality? In the orthodox NL treatment, the expectation of full-visibility rotational invariance can be framed through a few primary premises: (i) Since Heisenberg uncertainties preclude attribution of intrinsic properties to discrete quan- tum entities, entangled states must be treated as in an indeterminable superposition of all possible states during evolution to the measurement context. (ii) Dichotomic spin states, essential to both the wavefunction and the matrix operations, are assigned propensities such that on operation of the Pauli matrices, the entities are actualized with real vectors in a frame aligned with the discriminator reference frame. (iii) The matrix operations are assumed to represent a physical behavior leading, under experi- mental conditions, to actualization in alignment at the discriminators.

24 Ironically, the first premise sets up the entangled pair in a state from which vectorial infor- mation is excluded. Since correlated vectorial properties, revealed on measurement at space-like sep- arated stations, are essential to the outcome, the processes through which they do become available and aligned, should have raised all those concerns implicit in Einstein’s “incompleteness” argument. Extending the irony, Bell justified his use of the term “hidden variables” in the context of that argu- ment20. However, quantum uncertainties demand a superposition, subsequent application of the or- thonormal treatment provided a consistent resolution, and von Neumann’s formalization88 showed that no “hidden variables” were needed. Although Bell14 in his critique of von Neumann found logi- cal inconsistencies that allowed “hidden variables” in some contexts89, he also found that they did not apply to that QM treatment, and he therefore saw no reason to invoke them in that context. Instead, in a triple-irony, the inherent difficulties were transferred to local realistic theories. The “hidden varia- bles” were needed there only because in setting up his OL model, Bell had stripped the “entangled” state of its vectorial character. The vOL model now introduces a scenario comparing local and non-local perspectives in which both models are vectorial and compliant with seminal QM essentials. In my vOL interpreta- tion, all energy exchanges are necessarily local and quantized. When the vectorial information is car- ried by the photon, properties are intrinsic and explicit, and therefore cannot be “hidden”, and no conspiracies are needed to explain the outcomes simulated. On the other hand, the above premises necessarily leave the conventional NL model still as “incomplete” as it was when Einstein challenged Bohr1,10,85.

3. Indeterminacy and superposition of the entangled state A fundamental tenet of quantum mechanics is that all energy exchanges at that level are quantized. From a research career in photosynthetic mechanisms, it is obvious that at the molecular level interactions of photons with electrons in molecular orbitals involve local exchanges in which energy is conserved, the charge of the electron is inviolate, and photons are neutral. This is in con- trast with the electromagnetic nature of light inherent in Maxwell’s equations, and the interaction of light with harmonic oscillators through the fields of the wave. This distinction is discussed below, but the neutrality of photons makes mechanistic involvement through electromagnetic properties un- tenable, and I will therefore adopt the former perspective as my starting point.

The uncertainty principle and its application

25 Despite the extensive literature, I can see no reason to believe (in the context of entanglement experiments) that quantum uncertainties should exclude treatments in which discrete photons have intrinsic properties. In the conventional treatment, superposition is called for because uncertainties preclude assignment of definite properties to quantum entities. The treatment dates back to the period when the electronic orbital occupancies of the H-atom model were being sorted out, and the minimal ℏ uncertainty derived for electrons by Heisenberg (휎 휎 ≥ in the 3-D case) then precluded assign- ment of definite properties. The argument was applied in a scenario of simultaneous measurement of conjugate variables for position and momentum on a single electron. Can this approach be applied to photons? Although a similar minimal uncertainty has been suggested for photons90, the logic of this approach simply cannot apply. Since detection consumes the photon, it can occur only once; there’s no way to detect a photon twice, so the equivalent experiment could never be attempted. But that is not a problem in the entanglement context because we’re dealing with two different photons, interrogated through refraction, and separately detected at different stations. Measurement involves two distinct components, - interrogation and detection. From the ex- perimentalist’s perspective, although with photons, detection is a one-off event, it does allow one cer- tainty; a recording of the time and place of arrival of the photon. This here-and-now information is all that is directly available from detection, and leaves as a separate issue how to determine other photon properties. To access those, experiments have necessarily examined populations. In principle, the properties of the photons in a population can be determined when the source and pathway of evo- lution are known; a measurement can then be interpreted in terms of information available from the generating transition, and from interrogation during evolution to detection. Interrogation is of interest only when it does not consume the photon. The processes can be selective (transmission through a filter, prism or monochromator), reflective when mirrors are used, or refractive in, for example, use of lenses, HWPs or polarizers. Refractive events are loss-less, so a single photon usually experiences multiple refractive interactions in its path to detection. Useful information can be gleaned from anal- ysis, because, without changing the energy of the selected photons, the path is perturbed in time and/or in space. At a particular frequency (defining the ), refractive behavior probes the remaining property of the photon, the polarization vector. To engage, the vector of the photon must match a vector for electronic displacement, - along the polarizer axis in a polarizer. The proba- bility that a photon will excite a displacement will then be given by Malus’ law. The uncertainty here is in that classical probability term. Since elemental measurements are made on populations, such

26 statistical uncertainties are ironed out in the mean. Epistemologically, the essential point is that be- cause in the entanglement context, two separate photons are detected, and the refractive events are loss-less, quantized interrogation entails only the statistical uncertainty inherent in probabilities of Malus’ law at the elemental level. Bell saw the distinction between the views of Bohr and Einstein as between “…wavy quan- tum states on the one hand, and in Bohr's ‘classical terms’ on the other…” (see Preface81, and 91). From the , either a particulate or a wavy treatment could be used; choice of one must then have an equivalent expression in the other which gives the same result. In the entan- glement context, Bell seemed to equate Einstein’s view with the classical. As a consequence, with NL expectations required a treatment in their wavy guise, with the entangled pair in superposition, and with Bell’s model excluding vectorial properties, consideration of Einstein’s perspective was ex- cluded, even though it was the obvious particulate QM choice. Following Dirac92-94, advances over eight decades in (QFT) have dealt with the particle/wave duality in the wavy realm by coopting Maxwell’s electromagnetic wave treat- ment; different interpretations have led to many different models95 and a massive literature demand- ing a technical expertise beyond my skill-set. However, a simple conclusion requires no such exper- tise. Apart from the claim to have eliminated Einstein’s model, no experimental test has been sug- gested to eliminate any of the others. Since the BCHSH argument did not explicitly invoke Max- well’s approach, further discussion will be deferred to the concluding section. In the simulation, I im- plement the vectorial consequence of energy exchange by momentum transfer, and ambiguities from Maxwellian borrowings are simply bypassed. For entangled photon pairs, the complexities inherent in complete representation in Hilbert space have perhaps also been overhyped. The spatial complexity is restricted because the action vec- tor is orthogonal to the z-axis of propagation. This limits measurement of vectorial parameters to the x, y plane, which simplifies consideration to the unit circle (Fig. 2 and legend). In the conventional NL treatment, the spin states specified in the wave function are used in the matrix operation to gener- ate real vectors for the photons; since the only known vectors are those of the polarizers, actualiza- tion must necessarily depend on reference to the polarizer frame to implement any alignment. The mechanism for alignment itself is discussed extensively below. The photon vectors used in the unit circle projections are aligned, and could be seen as electric or momentum vectors, and either as com- ing from the complex plane, or as the intrinsic vectors of the vOL treatment. Observables returned on

27 taking squares have the same Malus’ law values in either treatment, so the only component of the treatment that discriminates is the implementation of alignment. Practice elsewhere in physics deals in photon populations, but without any prerequisite for such complexities. For example, in determining the history of our universe, modern cosmology in- vokes three simple notions. (i) Measurements on a uniform population can reveal common properties of the discrete photons making it up. (ii) Those properties are intrinsic and determined by the transi- tions in which the photons were generated. The spectra inform us of the local chemistry. The proper- ties may also be modulated by local fields, for example to generate polarization. (iii) Properties are, in principle, conserved on travel over space-like distances. Although frequencies are modified during their journey by relativistic effects, and gravitational refraction can distort the path, these features al- low interrogations that provide, in a measurement context, information about the source and its envi- ronment. Without these assumptions, we could not construct a history of the cosmos. Similar princi- ples underlie all spectroscopic applications, and my simulation of vOL photon populations is based on these same assumptions. If we recognize with EPR that each photon of the pair carries its own vectorial property, the idea of an ontic indeterminacy loses any meaning. I can see no simple model of the photon other than Einstein’s that is compatible with the quantized nature of its interactions, the time and length scales given by its frequency, and with rela- tivistic constraints. Uncertainties are statistical, - an epistemological challenge. A photon can only be detected once, but in the entanglement context we have two of them, and refractive interrogation is loss-less, so the logic of the uncertainties from measurement on a single entity cannot be taken to ne- cessitate a treatment starting with states in superposition.

4. What is known from the generating transition? Under all experimental protocols reported, useful information has been available about the transition generating the initial state. The correspondence principle8 would require that such infor- mation be considered under any QM treatment. Dirac94 suggests that indeterminacy and superposi- tion necessitate a probabilistic treatment, with an outcome that “…expresses itself through the proba- bility of a particular result…being intermediate between the probabilities of the original states…”. However, it is the itself that demands a probabilistic treatment. Probabilities give Malus’ law yields, so the result of a set of pairwise measurements of vectorial correlations should simply generate the statistical outcome expected from such properties, as it does in my simulation. Problems arise not in the vOL case, which, by recognizing discrete properties, allows a natural treatment, but in the NL case, where, contrary to the correspondence principle, vectorial properties are excluded. With

28 a dichotomic Bell-state, all that can then be represented is the pair-wise correlation of spin topologies (the phase difference, or potentialities) implicit in the spin quantum numbers. A mechanism generat- ing specific vectorial properties in the process of actualization must then be invoked to account for the outcome. (a) Cascade sources. With cascade sources, frequencies are known from spectral lines, experi- mental limits for vectorial correlations from conservation of angular momentum are well known; these determinate properties are used in design of protocols. When photons from arc-discharges are used to excite the atomic beam, since orientation of both the atom beam and the photon source are stochastic, orientation of the pairs after excitation must also be stochastic. Their stochastic nature has real consequences. Excitation of the atomic beam by a laser would lead to photoselection36,37. This was demon- strated by polarization detected along the orthogonal y-axis when the laser was polarization along the z axis of propagation36. Photo-selection along the z axis would excite a population of atoms with vec- tors in that direction, but it would be isotropic in the x, y plane of measurement. Experimentally, the singles-counts measured in that plane would then follow the stochastic pattern36, as expected from this analysis. (b) Sources generated by parametric down conversion. Laser excitation of PDC provides well- oriented photon pairs correlated through and angular momentum (phase- matching96); all properties are defined with respect to those of the pump laser54,80,96,97. The photon pairs separate into two populations, orthogonal in orientation, with determinate vectorial properties (H and V are real orientations in the pump reference frame); they are, in Bell’s usage14, prescribed. The behavior of one population is independent of measurement in the other18,80. The PDC output is clearly a state the “with prescribed values” of Bell’s caveat14. Should we not take Bell’s conclu- sion seriously, and worry about the adequacy of the NL approach in that context? PDC sources have been used in all recent photon-based experiments claimed to support the non-local perspective (cf.17,33,41,47,80,97), and knowledge of these determinate properties is exploited in design of experimental protocols. For example, in a seminal paper41 entangled pairs were generated in complementary cones by PDC in a type-II phase-matching BBO crystal excited at 351 nm. Pho- tons at 702 nm were selected from the two intersection points of overlapping H and V cones. Since the Bell state was HV/VH, the photons at either intersection were not pairwise correlated; for any pair, the entangled partner was in the other intersection. After further tweaking by insertion of half-

29 wave plates (‘state-preparation’), the mixed populations were sent to separate stations for polariza- tion analysis, and measurement. In another application54, compensating crystals were used to maximize the output of entan- gled photons from a double-crystal BBO type-1 source, - a pretty exercise in optical design. 80 In a type-II configuration using a different crystal for PDC (periodically poled KTiOPO4) , and pumped at ~400 nm, the co-linear cones at ~800 nm overlapped completely, but the contribu- tions of the two orthogonal orientations could be distinguished by polarizer rotations80. Entangled partners at the chosen frequency were in opposite halves of the overlap, so could be separated using mirrors and irises, allowing a much larger fraction of the population to be tested. A recent variant of this approach86 used the same crystal for PDC, but configured in a Sagnac interferometer. This configuration, discussed in detail in the SI (section 6 Bb, and Fig. SI_1A), was also used in entanglement experiments with the source located in a satellite to test communications loop-holes in58, and in another recent spectacular over-kill in the context of such loopholes59. As dis- cussed in the SI, some features of the Sagnac configuration are far from conventional. In particular, the signal and idler beams were fully polarized when projected to measurement stations, and the H and V photons at each station came from the two separate PDC processes, so neither station sees the mixture in ontic dichotomy implicit in the conventional NL treatment. The general point I want to emphasize is that both in cascade experiments and PDC-based protocols, detailed information is available from knowledge of the source. Manipulation of the “en- tangled” populations in state preparation is clearly based on exploitation of that information; on spe- cific determinate properties of the source, and on classical behavior at refractive elements. Uncer- tainty is introduced from a mixing of two separate populations for which at least some properties are determinate. With well-determined PDC sources, this generates a determinate dichotomic state quite different from the dichotomy by predicate of the orthonormal application. The appearance of indeter- minacy from mixing does not mean that intrinsic properties and their correlations are lost. When in- trinsic, photon properties would be retained, and photons would then behave at polarizers according to Malus’ law80 to generate my vOL outcome. Given that the initial state is determinate, two ontic changes would be required to generate NL expectations; one, after generation of the pair, to a state in superposition in which vectors are lost, and the other, the notoriously vague “reduction of the wave packet”35, to the determinable states, revealed on measurement, with vectors aligned with the polar- izer frame. Absent a mechanism, this is plain silly.

30 5. How is full-amplitude rotational invariance implemented under NL predicates? Two overlapping scenarios need to be examined, the first a mathematical model, and the sec- ond an attempt to relate the model to processes in the real world. The model. Bell used the conventional QM approach13,20 for entangled electron pairs, summa- rized in the matrix operation of eq. 2. In the BCHSH consensus, essentially the same QM approach was applied to photon pairs9,15. As shown by eq. 2, the matrix operations are treated as vectorial, in- volve photons and polarizers at both stations, and an implicit simultaneity of action. As discussed be- low, the spin correlations determine the potentialities (the phase difference), but these are then taken to also indicate propensities. In operation of the Pauli matrices, the propensities in effect determine that the photon pairs become aligned with the reference polarizer. If the photon was H, it will emerge in the ordinary ray, if V in the extraordinary ray, and the partner photon at station 2 will be actualized with the orthogonal vector implicit in the phase difference. The constraints of the matrices are then resolved as observables, - the Malus’ law yields. Shimony provides a useful summary19 of the above; in effect, the alignment of the photon frame with the fixed polarizer implemented in the matrix opera- tion is described “…by substituting the transmission axis of analyzer I for x and the direction perpen- dicular to both z and this transmission axis for y”, where x and y are H and V in the wavefunction equation. To attain the outcome claimed, the primary premises (Section 3) require auxiliary assumptions. 1) Application of Malus’ law requires vectors for photons and polarizers at both stations, so actual- ization would have to occur before the photons reach the polarizers. Since vectorial information for the photons is exclude in the superposition, vectors have to be provided in the process of actu- alization. The fixed polarizer is the only reference frame, and the difference from the variable po- larizer is referred to that zero. The potentialities of the spin states are represented by the binary matrix elements. In effect, the propensities take on a vectorial agency in the matrix operations, and then implement an alignment constrained so as follow Shimony’s prescription above, ac- counting for the outcomes claimed (see Section 6, 3) and 4) below). 2) In recent applications using PDC sources, both polarizers are fixed, and their discriminator func- tion is implemented by using HWPs to rotate the beams before they arrive at the polarizers. This requires a modified treatment to cover the HWP function, - actualization has to occur at or before the discriminating HWPs. However, since similar refractive components are used upstream in state preparation, the question of why actualization does not occur there becomes problematic. 3) The NL outcome depends only on the angle difference between polarizers, σ. Operation of the Pauli matrices implies simultaneous actualization at both stations; in the math this is no problem,

31 because both polarizers are engaged, but in the real world, the two stations are space-like sepa- rated, and information about polarizer settings is only available locally. The question of how sim- ultaneity is achieved then becomes problematic.

6. Relating the model to the real world The math may be elegant, but the physics is not. Several features are missing, mainly because the spatial element is ignored in the math but cannot be avoided in the physics. Firstly, when starting from a stochastic state, or with a misaligned ordered state, an input of work is required to generate the aligned state needed to explain full-visibility. Secondly, if the polarizers are to retain their natural function, photons must have real properties before they arrive there. Actualization in the ordered state would have to precede interaction with the polarizer. If HWPs are used to provide the discrimi- nator function, actualization must be before them. Thirdly, the stations are space-like separated; a value for the vector for the photon at the reference station can be known only at the instant of actual- ization, but that information is needed simultaneously at the other station to allow actualization of the second photon in appropriate correlation. Fourthly, in the mechanism suggested the spin states are assigned propensities with a causal role, but there is no obvious physical basis for this. Analysis of these anomalies reveals a common problem: the premises do not match the properties of the system under study. 1) There is no evidence for any intrinsic vectorial dichotomy in stochastic populations of electrons or photons. The conventional formalism requires an intrinsic dichotomy of spin states. Interpretations involving such states date back to the seminal work in which a beam of sil- ver atoms, each with an unpaired 5s electron, was analyzed using Stern-Gerlach magnets98. The atoms were aligned by the magnets into two well-defined populations; the partition of the atoms was determined by the spins of their electrons. This was interpreted as showing that electrons come in two different spin states, ↑ (up) and ↓ (down). The partitioning into well-defined popula- tions was then interpreted as showing an intrinsic dichotomy representing propensities for parti- tioning. The spin states became aligned with the field of the Stern-Gerlach magnet as determined by their propensities; the up electrons were expected to emerge in the upper population, and the down electrons in the lower population. However, the dichotomy implicit in this assignment can- not be taken as ontic, because, as fundamental particles, all electrons must be the same (cf.99). Expectation of such a dichotomy came from a misunderstanding. The dichotomy of the discrimi- nator function was interpreted as demonstrating an intrinsic dichotomy of the particles. This di-

32 chotomy has become deeply embedded in the standard orthonormal treatment, and the potentiali- ties implicit in the spin quantum numbers have been invested with the vectorial agency of pro- pensities. Although spin may be an intrinsic property of the electron, spin quantum numbers are not. Like all the electronic quantum numbers of the hydrogen atom, they are topological designators. In their atomic context, the topologies of orbital occupancy are defined by the lower quantum numbers, and the electrons in completed orbitals are paired. The spin quantum numbers, s = ±½, determine the relative orientation of the spin states, with potentialities (the phase difference) given by π/2s; the spin axes of a pair then have opposite vectors (↑ and ↓) so that their magnetic effects cancel. This makes sense in the context of electron pairing; they really are entangled through the work term expressed in the additional bond stability, and the orthonormal operations then give the ensemble result implicit in modern quantum chemistry. But such topological conse- quences cannot be in play with unpaired electrons. When unpaired, nuclear and electron spins can be explored experimentally using NMR or EPR (in this context electron paramagnetic resonance, not to be confused with the authors). When placed in a magnetic field, they are then found to distribute into separate populations fol-

lowing a Boltzmann distribution, = 푒푥푝 − , reflecting, for electrons, the energy differ- ence between up and down states. The fractional excess of the ndown population, which becomes more obvious as T is lowered, leads to a net absorption when the populations are flipped on illu- mination by radio or microwaves at resonance. The Boltzmann distribution in the magnetic field does not shows intrinsic dichotomy. If the spins were in an ontic dichotomy, they would distrib- ute evenly, and no NMR or EPR signal would be detectable. If electrons are not intrinsically dichotomic, the sorting has to be explained by an alternative mechanism, and a trivial one is available. The separation by the magnets into the two ordered spin populations reflects two different but overlapping processes in measurement. With atoms (and their 5s electrons) initially at stochastic orientation, spins would have partitioned with equal probability to either the ↑ or ↓ channel on encountering the magnetic field, based on simple pro- jections. They are ordered into distinct sets through an overlapping “active” role of the magnetic field, which provides a force to pull the spins (and their atoms) into two well-defined popula- tions, just as happens in the field of the NMR or EPR magnet. The observed distribution into well-defined populations does not reflect an intrinsic dichotomy in the incoming population, but its stochastic nature, the initial partitioning, and the active alignment.

33 2) The alignment of photons at a polarizer does not require dichotomy. The assumption of dichot- omy embedded in the treatment of electrons led to the standard orthonormal treatment, and was generically applied to the treatment of photons, but it is even more obviously inappropriate in that case. The observable properties, the polarization vector and the frequency, are both deter- mined by the generating transition. A polarization analyzer has refractive elements in orthogonal orientations, partitioning photons into H or V polarized populations, but, although this might sug- gest an intrinsic dichotomy in the photon population, 200 years of experimental application (cf.100,101) shows no justification for any such property for photons. In those cases where coupled transitions yield pairs in dichotomic spin states, the correlation is determinate, and well-explained by conservation laws. In entanglement experiments, the polarization vector of each partner is the critical property tested through refractive interactions. The alignment observed can be explained economically by the quantized nature of each interaction. When refraction occurs, the full action (hν) is expended in the displacement of an electron (the initial quantized event) over a timescale represented by 1/v. The lossless recoil returns a photon with the same action and with the vector of the reverse displacement. The displacement vectors available in the polarizer are aligned along the polarization axis, at angle φ (or orthogonal in the extraordinary path). With the photon simi- larly aligned, the displacement and its reverse have the same vector. If not aligned, the behavior 2 will reflect the Malus’ law probability (cos θ) that a photon with its momentum vector at λinit can

induce the transition along the vector φ, with θ = (φ – λinit). For the fraction given by this proba- bility (the remaining fraction goes to the orthogonal path), the photon is returned with a vector

reflecting the recoil, with λfinal = φ. Subsequent refractive engagements would involve this same vector for electron displacement, to give the polarized outcome determined on measurement. With an ordered population, at any particular polarizer orientation, all photons will contribute to a mean to give the Malus’ law yield; with a stochastic population the distribution will follow a 0.5cos2θ curve, centered on the polarizer vector (see Part A 5, (ii)). This explanation is local, re- alistic, and completely natural, is consistent with 200 years of experimental validation, and re- quires no intrinsic dichotomy. 3) The matrix operations do not describe a natural process leading to alignment. In protocols using photons to test Bell’s theorem, the three frames involved (that of the correlated photon pair, and those of the polarizers at two separate stations), are defined by space-like separated actions, - generation for the pair, and measurement at the stations. Evolution of the partner photons to measurement necessarily involves independent space-time histories. Orientation for a particular state should entail reference to its independent frame, because the frames can be (and, in many

34 protocols, are) rotated independently. For the photon pair, the frame is initially provided by the transitions generating the partners, but this information is discarded in the NL treatment. It may be mathematically elegant and convenient that the matrix operations actualize the vectors of the photon pair in the common frame of the fixed polarizer, but this is a mathematical device that in- cludes no physical mechanism for alignment, and no recognition of complications associated with the distribution of processes over space-like separated locations. 4) Quantum ‘magic’; identifying the Maxwellian demon. In operation of the Pauli matrices, the only vectors know are those of the polarizers. In the math, the matrix elements represent potential states for photons of a pair, the matrix operation involves both polarizers, and all processes occur together. However, the real actions are space-like separated, so neither station has access to all the information needed for the full operation. The expectation of rotational invariance depends on alignment of the photon frame with the fixed polarizer. Even if we accept the convention of an ontic dichotomy, the spin quantum numbers, s = ±1, carry no vectorial information, neither do the spin states, nominally H and V, derived from them. Only the spin states are represented in the wavefunction. The correlation through the phase difference may appear to be vectorial but could become so only by reference to a real vectorial frame. The only frame defined is that of the polar- izers. Nevertheless, the matrix operations lead to actualization of photons in alignment with that frame. This ordering of the state requires a work term, but none is available. In effect a Maxwel- lian demon is then needed to create order out of disorder. From the above, the action of the de- mon comes from the assignment of a vectorial agency to the spin states. In the math this works because the operation is vectorial, but the matrix elements are binary; at station 1, either an H photon represented by a 0 element in the matrix becomes a vector aligned with the reference po- larizer at 0, or a V photon is actualized orthogonal. In either case, the partner photon at station 2 is actualized in the complementary orientation. Then on rotation of the variable polarizer, the out- come is as shown in Fig. 2. But this is not natural; in effect, the propensities pre-determine the alignment. 5) The vectors required for projections at the variable polarizer (station 2) depend on instantaneous transfer of information about the partner photon actualized at station 1. As noted above, the ma- trix operation implements all processes in the frame of the reference polarizer, but without any recognition of problems associated with spatial separation. In real experiments, no information is held in common by the experimentalists at the two stations. The polarizer vectors are known, but only locally. The matrix operation leads to simultaneous actualization of photon vectors, both in the frame of the fixed polarizer, correlated through their phase difference, and aligned through

35 their propensities. However, the partner photons are necessarily at the two separate stations. If the reference polarizer is at station 1, the vectorial information becomes available there, but only at the instant of actualization. Actualization of the photon at station 2 with vectorial propensity ex- pected from the event at station 1 could not happen simultaneously, because the information needed at station 2 could not arrive in time. The absurdity would disappear if the superposition consisted of discrete photons carrying propensities as real intrinsic properties, - essentially as vectors in a frame pre-aligned with the reference polarizer, - but such properties are not allowed in the superposition. Only in the vOL model do photons have correlated real vectors in an explicit alignment as intrinsic properties. Then, nothing else is needed, because all the information ar- rives with the photon. Even if quantum entities were always to arrive in a superposition with a dichotomic distribu- tion of states correlated as suggested by the spin quantum numbers, the NL expectations would still require a real mechanism for alignment. There is no obvious justification either for a superposition, or for such a dichotomy, or the alignment. Since the alignment implemented by the demon is now ex- posed, the treatment looks to be inadequate on all counts.

7. What do the experimental protocols tell us about the processes needed to implement the NL expectations? In the experimental context, the entangled state must evolve in a spatiotemporal framework from its site of generation to measurement at space-like separated stations where real properties are actualized. As discussed above, the treatment is missing several important mechanistic processes for causal connection between events. Rather than belabor these shortcomings, consider a more general question, - that of the ‘…notoriously vague “reduction of the wave packet”…’ that Bell worried about (cf.35) but left unresolved. Actualization can only happen once, and if the alignment imple- mented in the matrix operations is to be credible, would have to occur simultaneously in both chan- nels. For each refractive event, the photon has to be there, fully represented by discrete properties. With PDC sources, HWPs are used both in “state preparation”, and to provide the discriminator func- tion, but this compounds the vagueness (cf.41,80,86) because the HWPs must, unless special effects are invoked, behave the same in both roles, but actualization can occur only once. For example, in the Sagnac interferometric configuration86, each photon passes through eight refractive interactions be- tween the pump laser and detector. Each of these implements a specific local function critical to state preparation. Unless actualization occurred at the first element encountered, the failure to actualize the pair on any subsequent engagement has to be explained away by a different specific mechanism

36 and auxiliary hypothesis for each. Actualization at the first element would solve this dilemma, but if that happened, the lifetime of the state in superposition would be limited to a short (ps) span between the pump laser and the first HWP encountered. The transient superposition would then make no dif- ference to the outcome because nothing else happens in that gap, and the “state preparation” subse- quent to the first encounter is then determinate. Superposition is for practical purposes redundant, and any quantum ‘magic’ a fantasy. A “QM simulation” mode can be called in the program, which uses a few “If…then…else” statements (see code) to implement Shimony’s prescription for the outcome at station 1; this is sup- plemented by appropriate auxiliary assumptions for the behavior at the second polarizer. (It is as- sumed that polarizer 1 defines the frame for actualization.) With any population of pairs in a dichot- omy of states, the program then generates curves showing the peculiar invariant behavior (blue sym- bols in the left frame of Figs. 2, 5). However, any suggestion that this is realistic would be greeted with well-deserved derision by the wider physics community. Any such process would have to in- voke a physically coherent mechanism, including a work term that could reorder the incoming flux, and synchronization of timing and differential function for the polarizers at the two stations. Since no appropriate terms are available, the routine would have to invoke a Maxwellian demon to implement the invariant behavior. However, as noted above, this is also necessary in the NL treatment. The ma- trix operations of the NL treatment generate the same outcome as the demon of the simulation, in ef- fect by implementing the same alignment, also without identifying appropriate work terms. Any such process would then necessarily be encumbered by the ‘tensions’ with relativity and the second law19. They exist because the processes postulated are unnatural.

8. In summary, the NL predicates are surprisingly flimsy (i) Although the all or none feature of quantized events must limit the precision of elemental measurements on a quantum entity, there is no reason to suppose that this excludes intrinsic proper- ties. The conventional justification for superposition in terms of simultaneous measurements of con- jugate properties on a single entity does not apply in pairwise measurements on separate partners in a population. A photon cannot be detected twice, but that does not matter because we have two of them, and only need to detect each once. Refractive interrogations return E in full, so no un- certainty invoking hν in a causal role can be justified. (ii) When determinate and separate H and V populations are generated by PDC, they cannot be in a superposition and are not indeterminate. This is particularly obvious in configurations using a

37 Sagnac interferometer (see SI, Section 3 ii) b), c)). In more conventional applications, the mixing of determinate populations from separate cones in state preparation modifies neither the discrete proper- ties of each photon of a pair, nor the correlation between partners directed to different separate sta- tions. (iii) The matrix operations through which alignment of the photon and the polarizer frames is im- plemented represent a mathematical device, not a real process. The alignment of frames depends on assignment to spin states of propensities, and their representation in the binary matrix elements which, on operation with reference of the fixed polarizer a zero, have vectorial consequences with no physical justification. The attribution of propensities is based on a misunderstanding of the relation between properties of the discriminators and the mechanism leading to partitioning of the outcome. (iv) The matrix operations require a dichotomic photon population, but there is no reason to sup- pose such dichotomy is ontic. Even with dichotomy, there is no reason to believe that propensities cause differential partitioning. With PDC sources, the dichotomy is determinate, reflecting real vec- tors that do cause differential partitioning, but that is fully expected under OL predicates. The para- phernalia of the NL treatment is needed only because vectors inherent in the generating transition are excluded in the superposition. (v) The ordering of a disordered photon population occurs without a work term, - the characteris- tic of a Maxwellian demon, whose machinations are now explained. The assumption that the propen- sities align the spin states implies a causal role which has no basis in physics. Even if it occurred, the process would still lack a work-term for alignment and a mechanism for simultaneous actualization. None of these features, - superposition, ontic dichotomic properties, alignment steered by propensities, - is justified. All are necessary for prediction of the NL expectations.

9. Experimental outcomes In comparing simulated with experimental outcomes, it is worth noting that the simulation is ideal: pairs are generated simultaneously, all photons are counted, and the coincidence window is vanishingly narrow (there are no “accidentals”, and corrections are unnecessary). This simplicity is useful but may mask features of consequence in interpretation of important functions. In particular, refractive behaviors are modeled in terms of linear polarization and empirical behavior, without any attempt to treat the complexities in the wavy realm implicit in phase delays or elliptical elements in- troduced by HWP/QWPs. With this caveat, the simulation is useful in addressing the question of how well the two hypotheses survive Popper’s test102. Does the behavior in the real-world match that ex- pected by NL or vOL models?

38 i) Experiments with cascade sources The early experiments with stochastic pairs from atomic cascades provide the most obvious challenge to local realism. The orthodox NL treatment was developed in this context and led to ex- pectation of the full-visibility rotational invariance, which, with a stochastic state, cannot be pre- dicted by OL models. What we would expect from a naïve perspective? The cascade occurs on decay from an ex- cited state (which can be generated through several different protocols) in two sequential transitions, distinguishable by their different energies22,36-38,40. Conservation laws determine that the photons fly off in opposite directions. In all reports, coincidences on pairwise measurement at different stations were maximal with aligned polarizers, demonstrating that partners always had the same orientation in the x, y plane. From knowledge of the source, the correlated kets might be represented as

|휃 휃 or |휃 휃 〉... |휃휃〉 or |휃 휃 〉, where each partner in a pair has the same orientation, and (for example, in a Ca-cascade) is either blue (4227 Å) or green (5513 Å). Experimental protocols are designed to take advantage of all these properties, but use of different filters at the two stations is at the expense a loss of half the coincidences. In a stochastic population, every ket must be at a different orientation, readily dealt with by simulation (as here) or analysis62,73 to give the LH2 curve. On the other hand, to match NL expectations, an alignment of each pair with the fixed polarizer would be needed to get full visibility. If this happened, all pairs would become aligned with the reference po- larizer, and the outcome would be polarized. No experimental result has shown this polarized behav- ior. To explain this, two conditions are needed, the ontic dichotomy of pairs in orthogonal orientation (nominally HH or VV), and actualization of each pair in alignment with the polarizer frame, inde- pendent of θ, but different for HH or VV. Can a detailed analysis of the experimental and analytical protocols distinguish these two possibilities? To avoid disrupting the flow of the argument, details have been relegated to the SI (see Section 6A), but the conclusions can be summarized as follows: (a) NL expectations with cascade experiments have been represented by an equation suggested by Freedman and Clauser22,

() = 0.25(휖 + 휖)(휖 + 휖) + 0.25(휖 − 휖)(휖 − 휖)퐹푐표푠2휎 (eq. 6). The result claimed (after adjusting for accidentals) was a rotationally invariant sinusoidal curve approaching the full amplitude (0 - 0.5 in this accounting). However, this is not expected from eq. 6. The first right-hand term (0.25(휖 + 휖)(휖 + 휖), RHT1) is independent of σ, so

39 could not contribute to the sinusoidal curve. The σ-dependent RHT2, which alone could generate the sinusoidal curve expected, would return a maximal value of 0.25. Overall, the equation would give a sinusoid of half-amplitude, offset by a constant value for RHT1. This is similar to the half- amplitude LR2 curve. Two differences, both previously noted by Thompson62,73, are in the ampli- tude (0.5 claimed, 0.25 expected) and the lack of offset expected from RHT1. The latter could be accounted for by cancellation on subtraction of accidentals, but there’s no accounting for the for- mer, and there are several other troubling features, as discussed in the SI (Section 3, i) a)). (b) Since the source is undoubtedly initially stochastic, to generate the NL expectation, each pair would have to become alignment with the reference polarizer, thus ordering the population. Can- cellation of accidentals would make RHT1 redundant, and a rescaled equation,

() = 0.5(휖 − 휖)(휖 − 휖)퐹푐표푠2휎 (eq. 7) would then be needed to account for the full amplitude claimed. This is, in effect, what the matrix operations promise. However, this would entail all the problems of the conventional NL treatment dealt with above, including missing work terms and lack of mechanism. (c) In a previous detailed analysis, Caroline Thompson62,73 had noted that the stochastic cascade sources would be expected to show the half-amplitude outcome (rather than the full-amplitude claimed), and she analyzed the data from Aspect’s thesis to demonstrate that this results could be naturally obtained by applying a different subtraction of accidentals. She also noted that a simple integration of pairwise expectations would result in a constant offset of the curve, not seen in re- sults reported. The LR2 curve of the simulation (Fig. 6) demonstrates the same result as her analy- sis. It is centered at the middle of the amplitude scale, and consequently is offset from the mini- mum and maximal values by 0.25 of the scale. Since the simulation has no “accidentals”, these offsets must be explained solely though the Malus’ law behavior (see SI, Section 3 i)). The offsets are expected from the stochastic distribution of pair orientations (Section 6, 2); with HV/VH popu- lations, practically all pairs will generate some coincidences even when the polarizers are nearly aligned, and some will fail to generate coincidences when they are nearly orthogonal. Integration over a population will then reflect the cancellations associated with these losses. Between these extremes, the simulation follows the half-amplitude cos2σ curve of eq. 6. The green curves in Fig. 6, calculated from orientations of a random pair sampled from each population, map out an enve- lope of possible Malus’ law outcome curves, which covers the full scale. The LR2 curve (which fits the simulated points) passes through the center of the envelope. The envelope is always sym-

40 metrical about the LR2 curve, indicating that the cancellations distribute symmetrically as ex- pected. Any vectorial approach compliant with second law and relativistic constraints would gen- erate the LR2 outcome, including the offsets, if the stochastic nature of the source was considered. (d) Fry and Thompson36 excited a 200Hg cascade using a laser polarized along the axis of propa- gation (the ±z-axes). The singles counts measured in the x,y-plane orthogonal to that axis where independent of polarizer orientation, - the rotationally independent behavior expected of the sin- gles counts at each station from the stochastic state or from a dichotomic distribution. However, strong polarization in counts along an axis (the -y-axis) orthogonal to the laser polarization was detected in the x,z-plane. If the populations were intrinsically dichotomic, the photons emitted along the orthogonal axis would not be polarized. (e) As noted above, if actualization on measurement yields pairs in simple matched orientation, then the lack of polarization in the pairwise outcome shows that they do not become aligned with the reference polarizer as might be expected under NL predicates. (f) The lifetime of the intermediate state means that the first photon will be displaced in time by ~3 ns, or ~1 m in space from the second. Since this distance is long compared to the wavelength, neither superposition nor interference effects could be justified except by ancillary hypotheses. The program includes a simulation of the Freedman and Clauser (FC) count. As with the other counting protocols, this produces the LR2 outcome simply by accumulating coincidences from pairwise counts reflecting Malus’ law behavior. Fig. 7 compares the curves for an HH/VV population, either with pairs both ordered and in frame with the fixed polarizer (circles), or at random orientation (diamonds), generated either by simulation of the FC count (cyan), or by a conventional coincidence count scaled to cos2σ (red). The curves generated using the FC count reflect the normalization proto- col above, and all are half the amplitude of those from coincidence counts, but otherwise show the same features. It is perhaps a little droll that, had the Freedman and Clauser result followed the expecta- tion of their equation, the outcome would have shown the half-amplitude LR2 curve. This would have given a limit of SQM≤1.414 when scaled to the conventional ±2. Since the OL model they expected was Bell’s zigzag showing a full amplitude curve and SOL≤2 at canonical angle differ- ence, this would have engendered an entirely different discussion of the “inequalities”. ii) Experiments with PDC sources.

41 Populations generated by PDC have correlated pairs in two separate orthogonally polarized output cones (H and V with respect to the pump laser as reference). In the vOL model, these proper- ties would be sufficient to account for all results except the full visibility of rotational invariance. In view of the inadequacies already covered, some skepticism might be appropriate. Have full visibility and rotational invariance both been consistently observed under all protocols? An alternative expla- nation for full visibility is that the alignments needed to generate NL expectations were established by refractive manipulations during state preparation. There are ambiguities in most descriptions of protocols17,33,41,54,80,86 that make it uncertain which explanation applies. In many cases, outcomes that match those reported can be readily generated in my simula- tion. Some examples are shown in Fig. 8 and others are discussed at greater length in the SI (see SI, Section 3 ii) and Figs. SI-2A,B). The analysis there suggests the following conclusions. a) The photons reaching the measurement stations are not in a superposition of states. The ever- increasing technical ingenuity of the experimentalists in design of optical configurations for test- ing the entangled states requires specific and detailed information about the determinate proper- ties of photon pairs generated by PDC, and explicit processes for refractive interactions used in state preparation. These populations are not in the indeterminate superposition essential to the premises of the NL treatment. Especially in the Sagnac interferometer-based systems, where all photons detected arrive at the measurement station in fully polarized beams, the notion of an in- determinate superposition is simply untenable (see SI, Section 6B, (iii)). b) Simultaneity of actualization is not needed to get the NL result. Whatever process is involved, it is not adequately described by the matrix operations. In recent tests, the communication loop- hole has been “closed” by testing entanglement at stations well-separated from each other and from the source. In several of these, the distances between source and the two stations differ sig- nificantly56,58,59,103. As a consequence, entangled photons arrived at their measurement stations at different times, and in several cases the difference in travel times was substantially greater than the coincidence window width in the ns range. Reconciliation of this difference could be achieved only by time-tagging photon arrival at each station (using coordinated atom clocks), and looking for the coincidence peak on sliding the data set at one station with respect to the other. For example, in Ursin et al.103, the source and one measurement station were located on La Palma, and the other measuring station was on Tenerife, and the time difference between detec- tion at the two stations was 487 μs. When data sets were centered at that time-difference, the co- incidence window was 0.8 ns. In Yin et al.58, using a Sagnac-configured source in a satellite, the

42 travel distance to measuring stations separated by 1200 km varied between 500 and 2000 km, and experiments required a sophisticated “…high-precision acquiring, pointing, and tracking (APT) technology…” to coordinate the temporal differences between different stations and the satellite source. This was achieved by transmission of additional information about orientation (via polari- zation of a separate supplementary laser) and about distances from the source (via a separate pulsed laser and GPS-synchronized atom clocks). This additional information was necessarily full determinate. As in103, time-tagged data sets were aligned by sliding to find the coincidence point. In Handsteiner et al.59, also using the Sagnac configuration, the buildings housing the re- ceiver stations were at significantly different distances (~500 m) from the source, and coinci- dences were again found by sliding time-tagged data. Clearly, in these experiments there is no question of the simultaneity implicit in the matrix operations. c) Closure of the communication loophole does not exclude local realism. In experiments by Weihs et al.17 and by Handsteiner et al.59, closure of this loophole has been used to test for con- spiracy in setting discriminators. Switching of discriminator settings between fixed values ex- ploring canonic angles differences, was triggered by a random number generators in17, or by the random arrival of photons from separate cosmic light sources in59. Simulation (see Fig. 8B) showed that the outcome in17 could be readily explained within vOL constraints, and thus sup- ports local realism (SI, section 3 ii), e)); in59, the settings themselves were only briefly described, but with ambiguities similar to those discussed elsewhere in this paper. Unless the ambiguities are clarified, these experiments do not exclude local realism. d) The NL outcome does not depend on each station receiving a state in ontic dichotomy. In the Sagnac interferometric configuration (Fig. SI_1A), all beams reaching the measurement station are polarized, and the results reported can be accounted for in the vOL framework simply on that basis (see SI 3 ii), b), c)). For each PDC process, the signal and idler beams go to different detec- tors so that the beams arriving at each station come from different PDC processes. There can be no question of a superposition of states. The mean singles-counts at each detector are 0.5, which might suggest a dichotomic superposition of states, but what the simulation shows is that it is a consequence of simple subtraction of polarized beams in a well-determined but out-of-phase cor- relation (Fig. SI_1B-D). Since the H and V beams arriving at each station involves different PDC processes, there is no way to describe the behavior through the conventional NL approach (see SI 3 ii) b)). e) No feature of the NL treatment is needed to explain the behavior reported. The alignment of frames in the matrix operation depends on actualization with reference to the fixed polarizer.

43 With PDC sources, the special rotational invariance could then be trivially demonstrated: - for any protocol demonstrating the full amplitude sinusoid, rotate the reference polarizer and repeat the experiment, without changing other settings. If the outcome is the same at any setting of the reference polarizer, the case is made. In all experiments reported based on PDC sources, a much more complex procedure, inspired by the canonical settings, has been used. In all of these proto- cols, the polarizers have been fixed, and discrimination has been implemented through beam ro- tations using HWPs or EOMs. The question of where actualization occurs in such protocols has been addressed in Section 7 above. Since Malus’ law behavior is, in all cases, invoked at the po- larizers, photons must be actualized with appropriate orientations before they get to the discrimi- nating elements. However, this does not fix where actualization occurs because HWPs and EOMs both implement beam rotations through simple refractive processes; if actualization is need there, it would also be needed in any refractive event further upstream. Despite the vague- ness of interpretation, the manipulations in ‘state-preparation’ are usually quite detailed and spe- cific, and clearly depend on knowledge of real properties. If all protocols start with the determi- nate H and V orientations of the output from PDC, and subsequent manipulations involve well- defined refractive behaviors, they would give the same outcome whether or not entangled state had been in superposition before the first refractive element. Actualization is needed only if the pairs were initially in superposition; if real states are involved, then none of the algebraic para- phernalia of the standard NL approach is needed. What is always needed for full visibility rota- tional invariance is the alignment of frames, always explicit in vOL treatments. f) How believable are the outcomes claimed to exclude local realistic interpretations? As an ex- ample of ambiguity in description of protocols, in recent reports from the Wong lab 80,86 either using a conventional PDC protocol, or one designed around a Sagnac configuration58,86, settings for the polarization analyzer in channel B were reported as 0o, 45o, 90o, or 135o. Implicitly, these represent beam rotations implemented by the HWP. Then, values for Bell’s SNL parameter con- sistent with full-visibility at all settings were reported when the dependence on angle difference, σ, was explored on rotation of the discriminating HWP in channel A. Since the polarizer in chan- nel A was set to 0o, it might be assumed that the polarizer at B was also set to 0o. However, the 45o or 135o rotation of the beam in channel B would then result in distribution of photons with equal probability to H and V channels of the analyzer. This would generate a zero-visibility out- come curve (open squares in Fig. 2, and similar in other Figs) because in pairwise comparison the H or V character would be randomly assigned. The full-visibility result claimed would then con-

44 tradict expectations from standard physics. The paradox could be resolved by recognizing an am- biguity in the description; the published account refers to settings for the polarization analyzer without explicit recognition of the two separate components. If the polarizer in the reference channel (B) is not set at 0o (as implied), but at the same angle as that to which the beam is ro- tated, the outcome is vOL compliant. This interpretation is not in contradiction with the protocol as reported, and simulation of this configuration then gives the same set of phase-shifted LR0 curves as reported (Fig. 8A, curves 2, 4, and Fig. SI_2). Alternatively, the same experimental setup could have been used to implement a slightly dif- ferent configuration, as in Weihs et al.17, in which there is no ambiguity. Both the frame shift and switching between rotations to explore changes in polarizer difference were made at the variable discriminator. This configuration (see Fig. 8B) leads, under vOL constraints, to the full-visibility outcome curves shown. Similar interpretations can be applied both to conventional protocols based on overlapping of cones, and to Sagnac interferometric implementations (see Fig. 8B and Fig. SI-2). In conclusion, since similar ambiguities in description of protocols are apparent in most re- ports using PDC sources, the take-home message from this section is that, within the limits of ambiguity, the vOL model can probably account naturally for all outcomes claiming to exclude local realism, and without contravention of standard laws.

10. Closure of communication loophole does not exclude local realism. Several protocols17,33,38,43,47,56,58,104, including some of the above, have been set up to elimi- nate loopholes involving communication between stations, generally by excluding subluminal trans- fer of information (see Sections 3, 4, and 6 of SI). Many of these involved impressive engineering feats, including measurements using astronomical telescopes at stations separated geographically by large distances (144 km in103), or with the PDC source in a satellite, with ground-based measurement stations separated by 1,600 km in58, or by rapid switching between specific analyzer settings or by beam rotations with photons in flight43,46,56,103,105. Demonstration that closure of communication loop- holes did not affect the outcome have been claimed to exclude local but allow non-local models. It needs to be re-emphasized that, although these experiments would eliminate models de- pending on subluminal communication between stations, they are of no interest with respect to Ein- stein’s model as represented in the vOL simulation. When photons carry intrinsic properties, vectorial information arrives with the photon. Since correlation between partners is conserved in all models,

45 no other communication is needed, and a natural behavior at the polarizers can account for the obser- vations reported. On the other hand, superposition, indeterminacy, and actualization in alignment on collapse of the wavefunction, require assumptions that both conflicted with the second law, and need superluminal communication between stations. These inconsistencies cannot be shrugged away. If the speed of light is the limit, closure of communication loopholes would, under conventional NL predicates, eliminate all such non-local hypotheses from serious consideration. The assertion that transfer of information is unconstrained by laws pertaining to the rest of physics18,19,48 might be thought to rescue the NL approach, but this lands one clearly on the metaphysical side of a useful de- marcation (cf.106, and Section 3 of the SI). The vOL model is the only one for which expectations are impervious to closure of communication loopholes.

11. Final conclusions The Bell’s theorem community has fostered a model that, by excluding consideration of vec- tors intrinsic to the initial state, by using spin quantum numbers to justify dichotomic states appropri- ate to orthonormal assumptions, by conferring causal properties to those states that, on implementing a mathematical device, generate vectors in the alignment needed, then predicts outcomes in contra- diction with standard constraints that govern the rest of science. If outcomes consistent with this model had been convincingly demonstrated, science would indeed be on an exciting path. However, for each of the reports examined above, the simulation demonstrates a simple explanation that is bet- ter aligned with foundational precepts, compliant with local realistic constraints and classical refrac- tive behavior, and is, within ambiguities, consistent with the published account. It is not obvious that the predicates of the conventional NL view are essential to a demonstration of results claimed as jus- tifying non-local interpretations, so re-examination is surely justified. The response from my colleagues has been varied, but can be paraphrased as follows (cf.107): “The many different interpretations do suggest that no QM explanation is entirely satisfactory, but they all have as a common feature a non-local perspective, and in that context, carefully designed ex- periments routinely show outcomes consistent with non-locality. The results are well understood, all the loopholes have been addressed, and there is no satisfactory mechanism consistent with local real- ism. That’s where the adventure in physics now lies; things really are spooky, but the paradoxes point the path to a future understanding”.

46 Before such a path can be found, some agreement is needed on where the paradoxes lie. Un- fortunately, the literature belies the assurance of my colleagues that the problems are well under- stood. Instead, the ad hoc nature of the predicates, and the determinacy issues are ignored, and the conflicts with relativity and the second law are either glossed over, glorified, or cloaked in metaphys- ics (cf. 108-113). The appearance of obfuscation is compounded by invocation of Bell’s limit of ≤2 as excluding local realism, - the main justification in many recent reports33,44,47,56,58,59,103. The vOL model, though an obvious alternative, does not seem to have been seriously considered, likely be- cause it falls outside the box. This leaves the community open to the accusation that it is acquiescing in an essentially metaphysical view of physics and philosophy23,109,110,114,115. Until the resulting con- fusion is cleared up, the “kind but bewildered listener”84 might legitimately retort that the line be- tween sophistication and sophistry seems thinly stretched. I cannot claim that explanations offered by my vOL model are exclusive; the results obtained under protocols cited herein were all interpreted as showing non-locality. With clarification, the in- formation already available would no doubt be sufficient to distinguish between these perspectives, so perhaps this essay will never be published. However, for each published report examined, the al- ternative vOL explanations are consistent with the protocol described and with local realistic con- straints. In general, and mindful of Einstein’s caveat (“…but not simpler”), Occam’s razor favors the vOL model, because, despite its naivety, it accounts for more with greater economy than the conven- tional NL one. Since that treatment is inadequate, I am open to any demonstration that an alternative NL treatment can without ambiguity fully account for the experimental observations while excluding local realism. has obviously advanced since the Bohr-Einstein controversy and its res- urrection in Bell’s theorem; indeed, Wikipedia currently has some 14 different interpretations of the nature of quantum reality, the central question in the entanglement debate. There are two clear con- clusions. Firstly, that the one model that has been tested, Einstein’s, should not have been excluded; secondly, as noted earlier, that the lack of a consensus reflects the fact that no experimental test has yet been devised to exclude any of the others. This state of play is obviously unsatisfactory from a philosophical perspective102. In developments following Dirac, Maxwell’s view of light as consisting “…of a coupled oscillating electric field and magnetic field which are always perpendicular …”116 has been co- opted as appropriate for treatment of photons. As the wave propagates, the fields interact with “harmonic oscillators”, mostly contributed by electrons in the medium through which the light

47 passes. A colleague has suggested that “action at a distance” is implicit in invocation of Maxwell’s fields in the wavy realm and can explain the simultaneous actualization at space-like separated loca- tions. Does this help the argument? That neutrality of the photon is problematic has already been noted, - they cannot interact through coulombic forces. When first presented, Maxwell’s treatment provided a spectacular synthesis, - a thing of beauty uniting all theories of the behavior of light through representation of their wavy nature, but required transmission through a medium in which the fields were en- gaged. The first blow to this edifice was the Michelson-Morley experiment that showed no effect on the speed of light measured when the orientation of the apparatus was changed with respect to earth’s trajectory through space. This was interpreted as showing that no such medium could be detected. Then Planck developed quantum theory requiring quantized interaction, and somewhat inconveniently, this was extended by Einstein in explaining the from a par- ticulate perspective. These impediments have been brushed aside by invoking particle/wave du- ality. In dealing with interactions between photons and electrons in quantum electrodynamics (QED), the probability expressions are essentially the solutions of Dirac’s equation92,93 for the behavior of the electron's , and Maxwell's equations for the behavior of the photon's probability amplitude. In interpretations through Feynman’s path integral approach, probability amplitudes are represented in Feynman diagrams. Although Feynman elegantly sim- plified the mathematical underpinnings117, the formal equations are highly technical, and, in view of my limited math skills, I must leave any depth in analysis to the professionals. But from Feyn- man’s account pathways are quantized, and to approximate the effective path, the integrals must include probabilities for all paths. Interference effects then come into play because phase delays from differences in the length and refractive index of the paths, and from any different frequency for the photons117, then tend to cancel contributions from paths more convoluted. Feynman showed that the path integral approach was equivalent to the Schrödinger equation applied to a stream of photons propagating in a wave (99,118, and see A. Zee in Introduction to 117). In empha- sizing that the treatment must embrace the particulate nature of quantized events, and that expec- tations must therefore be expressed probabilistically, Feynman found a brilliant way to make Maxwell’s electromagnetic treatment work, but the hybrid approach is ultimately futile because it

48 cannot readily be adapted to processes at the elemental level that necessary involve neutral pho- tons. In the wavy domain, the medium needed to support the propagation of a light wave is nowadays supplied through interpretations of the vacuum state as “… not truly empty but instead contains fleeting electromagnetic and particles that pop into and out of existence…”119, allowing particle/antiparticle pairs to fleetingly spring into reality on interaction with the fields of the passing photon wave. However, in light of the Michelson-Morley result, these interactions cannot involve any reaction time. Since quantized interactions involve engagements of their ac- tion that are frequency dependent, and necessarily linked to time, something else has to be in- voked to make this realistic. Mechanisms involving electron-positron annihilation generating 2 gamma rays photons, or a high-energy photon generating an electron-positron pair, are well known, but these don’t fit the bill; whatever happens has to be independent of frequency. Per- haps vacuum bubbles are involved120: “…Such charged pairs act as an electric dipole. In the presence of an electric field, e.g., the electromagnetic field around an electron, these particle–an- tiparticle pairs reposition themselves, thus partially counteracting the field (a partial screening effect, a dielectric effect)… This reorientation of the short-lived particle-antiparticle pairs is re- ferred to as vacuum polarization…”. Replace “around an electron” with “of the light wave”, and we have a suitable medium. However, such interactions are said to “…have no measurable im- pact on any process…”120; if this is so, the polarization through repositioning can involve no pro- cess engaging the known properties of photons, and is therefore immune to measurement. This is then just another hypothesis that cannot be tested. A more natural approach to dealing with photon neutrality would have to start by recog- nizing that the electromagnetic perspective is necessarily inappropriate. An alternative is to in- voke quantized transfer of momentum at the particulate level, where electromagnetic properties are not needed. At the elemental level, each path is defined by specific quantized refractive events, but all outcomes still reflect classical parameters (cf.121). Bialynicki-Birula 122 suggested a photon wave function in momentum representation which allows quantized interactions of neu- tral photons with electrons to be represented at that level. A more recent paper shows that the outcome is quite compatible with those of QFT approaches via Maxwell’s equations 123. In on- line discussion, “Maxwell”, an anonymous commentator124, finds a hybrid treatment in which a

49 wavy expression is developed following a form similar to those of QED, and a particulate ex- pression in which the total momentum of the photon is applied at the quantum level along the po- larization vector. Photons appear to be electromagnetic not because of any electrical or magnetic proper- ties, but because they are generated by, and their energy is used in, displacements of mass. In EPR and H-NMR, the flipping of spins requires photons of energies that reflect the ~1,800 dif- ference in mass. In the UV/VIS/NIR range, the most easily displaced mass, the electron, is charged, and the apparent electromagnetic behavior of photons then simply reflects this fact. In generation, the transition of an electron to a lower energy involves a spatial displacement that yields a photon with energy, E = hv, equal to the energy lost by the electron. The photon has a momentum vector along which action can be applied (the electric vector in the classical descrip- tion), which is determined by the vector of the displacement generating it. The photon can be thought of as a wave packet occupying λ = v/c of space-time, which propagates through a me- dium at c/n with overall momentum p = E(n/c) (where n is the index of refraction and c the speed of light in vacuo). When a photon transfers its energy to an electron, the momentum vector must match an available displacement vector for the electron, the energy involved in displace- ment must match hv, and the timescale of displacement must match 1/v. Events can be absorptive or refractive. If a higher-energy orbital (atomic or molecular) is unoccupied and the photon en- ergy matches the energy difference, displacement will lead to absorption. On the other hand, if the electron has nowhere to go, it has little option but to recoil in a loss-less refractive event re- turning the energy as a new photon. Whether justified in terms of momentum vector or the elec- tric field vector, projections leading to a Malus’ law outcome will be the same. A neutral photon does not require any medium to propagate through space. Passage of a photon through a refractive medium will involve sequential refractive events involving transfer of momentum and its return, calculable from refractive index, path-length, frequency, etc. Many potential paths will be explored on passage of a population. The question as to how phase-differ- ences are conserved if a process involves many sequential events and paths, was solved through the path integral approach in Feynman’s treatment, an approach that would work just as well in momentum transfer. Nothing in this description should be controversial. In contrast, the observer- dependent non-local reality is weakly founded, in conflict with relativity and conservation laws, is open to metaphysical exploitation, and could be abandoned for any of those failings.

50 My initial interest in the entanglement debate arose from a concern that invocation of such a blemished philosophy undermines the scientific enterprise; if relativity and the second law are dis- posable, what can be held dear? Perhaps what is needed is not new science, but a fresh look at the old science. Maybe a reformulation of wavy explanations to comply with locality constraints applicable to the particulate nature of the neutral photon and quantized nature of its interactions could be pro- ductive. From the wider perspective, a demonstration that Einstein’s local realism can account for most behaviors previously taken as justifying non-locality should be welcome, and its simplicity makes it a more attractive point of departure. In the SI (Section 4), I explore a number of cases in which the perspective of photons as momentum carriers leads to simple particulate explanations for problems, in particular interference effects, usually considered in the wavy domain. In a somewhat speculative discussion, I extend this viewpoint to cosmological controversies. This simple perspective might remove tensions, clean up the obfuscation, eliminate justifica- tion for metaphysical speculation, and perhaps even put natural philosophy back on a coherent track. If nothing else, my critique could focus the target better. The crutches of the entanglement edifice are clearly wonky; kick them away and perhaps a miracle might happen!

Acknowledgements. The biologist claiming to understand entanglement is a standing joke among quantum physicists, so I first wish to thank Mike Weisman, Martin Gruebele, Greg Engel, John Miller, and Paul Kwiat for their patience in helping me to understand some of the issues in early discussions. I also thank Thomas Angelidis for correspondence relating to his OL model. I acknowledge with gratitude a useful correspondence with Sean Barrett before his tragic death in 2012, and ongoing correspondences with Arto Annila and Richard Gill that have help me to appreci- ate quantum field theories, and other criticisms of Bell’s approach, respectively. I am particularly grateful to Tony Leggett for his permission to cite the distinction noted in his cases 1 and 2, and for discussion, correspondence, and criticisms that have corrected misconceptions and deepened my ap- preciation of the subtleties of the Bell’s theorem debate.

51 Figure Legends Figure 1. Schematic of a typical experimental set-up for an entanglement experiment12,19. A la- ser polarized at 0o (H) excites a BBO crystal, generating a population of pairs of orthogonally corre- lated photons through parametric down conversion. A filter at 702 nm (half the frequency of the 351 nm excitation wavelength) selects pairs of photons of the same color. These are emitted in ordinary and extraordinary cones, one H polarized, and the other V polarized. The cones overlap at two inter- sections where H and V photons are mixed. For each pair, the ‘entangled’ partner is in different over- laps so that correlated partners emerge in separate beams sampled at the overlaps18,41, with equal but random allocation of HV and VH pairs to the two paths. Each beam then has a random but equal mix of H and V photons, with complementary HV or VH pairs in the two beams. The photons are chan- neled, for example through optical fibers, to space-like separated measurement stations where polari- zation analyzers sort each photon by its polarization, with probabilities according to Malus’ law. In the experimental situation, four detectors, one each in the ordinary and extraordinary rays of polari- zation analyzers at the two stations (elemental outcomes Q, R, S, and T at the four output rays are equivalent to elemental probabilities: p1(λ, α) ≡ Q; p1(λ, α') ≡ R; p2(λ, β) ≡ S; p2(λ, β') ≡ T)), are used to measure the photons. Entangled partners are identified through the temporal coincidence. Photons are analyzed pair by pair at particular orientations of the polarizers, and the mean is taken over a sig- nificant population. The counting is handled by suitable electronics, involving timers, coincidence counters, etc., or equivalent recording procedures. Additional refractive elements (HWP0, HWP1, HWP2)) ‘prepare’ the entangled pair before measurement (others not shown may also be included, see text). In most recent reports, discrimination is implemented by rotation of the photon frame using HWPs P1 and P2, while the polarizers are static. The simulation can model this set-up by using HWP1 for P1 at station 1, or HWP2 for P2 at station 2, with the polarizer fixed. However, to keep the program simple, the default mode is to use the polarizers themselves for discrimination, with the var- iable polarizer at station two, and function implemented by rotation of the polarizer. HWP1 and HWP2 can then be used for “state preparation”. The simulation assumes classical properties for re- fractive components, and perfect optical elements so that no photon is lost to imperfection.

Figure 2. Typical outcomes from vOL or QM simulation, and a unit circle used to illustrate vector projections for aligned vOL or QM model expectations. Left: A Run3 cycle generates points at canonical polarizer settings. A vOL simulation using an ordered HV/VH population aligned with polarizer 1 at 0o (LR0 curve, ●); a stochastic population at the same settings (LR2 curve, △); or

52 an ordered HV/VH population, but with polarizer 1 at 45o (zero visibility curve, ◻). Curves gener- ated in “QM simulation” mode with a stochastic population (◻), or with HV/VH photons, but with polarizer 1 at 45o (△), or with ordered pairs at any orientation, fall on the QM curve (blue). Right: the unit circle representation. All vector operations can be represented without algebraic sophistica- tion. The vectors of the polarization analyzer at station 1 are at 0o (for the ordinary ray) and 90o (for the extraordinary ray). Under vOL premises, with an aligned population the vectors for an orthogonal HV (or VH) photon pair overlay these, with orientation of photon 1 at 0o if H (yellow), or at 90o (pur- ple) if V. Under NL treatments, the dichotomic spin propensities125 determine that the matrix opera- tions align the photon frame with the polarizer frame to give this same alignment. The black diagonal o lines are the unit vectors for 25 angle settings of the variable polarizer 2 (θ2) spaced at 7.5 intervals, restricted to one hemisphere between ±90o for clarity. These settings were varied with respect to po- o larizer 1 oriented at θ1= 0 , so that σ = θ2 - θ1 = θ2. At station 1, orientation of the H photon always aligns with polarizer 1, so that projection of polarizer 1 onto the two photon vectors would give 1 and 0 as values for cosσ and sinσ in the ordinary ray, with orthogonal projections in the extraordinary ray. At station 2, the red and magenta perpendiculars are projections of the polarizer vector onto the two photon vectors using the 25 values for σ, given by the intercepts at the circumference. (The projec- tions for the orthogonal polarizations in the extraordinary ray are not shown; yields are implicit in the remainders from yields given by the ordinary ray projections.). The text boxes show (top) values re- turned on clicking at an intercept (here at 22.66o (accuracy is limited by pixilation of the pointer posi- tion), for cosσ (left) and sinσ (right) (the potentialities or probability amplitudes), and (bottom) cos2σ (left) and sin2σ (right) (the Malus’ law yields, - observables or probability densities). For each polar- izer 2 vector, θ2 = σ, the intercept at the circumference gives projections whose values on taking squares give Malus’ law yields. The relation between these observables (cos2σ values) and the poten- tialities (cosσ) derives from representation in the complex plane, where the possible results from vec- tor projections are ±cosσ, falling in different quadrants, and potentialities in all quadrants are ex- plored in Hilbert space. In effect, since the only ‘moving part’ is the variable polarizer, the projec- ⃗ ⃗ tions at particular settings then represent the values from Bell’s analysis, (휑⃗ ∙ 훼⃗ 휑⃗ ∙ 훽)=−훼⃗ ∙ 훽 = −cosσ. The total probability for detection of a photon, cos2σ + sin2σ = 1.0, and the difference cos2σ - sin2σ = cos2σ, is shown for σ = 22.66o. The present implementation is limited to representation of the aligned situation. Under QM expectations, the dichotomic propensities are taken as determining that actualization is in alignment with the polarizer frame (see Part B, Section 6 for discussion of the va- lidity of the assumption). For a vOL population, the aligned configuration shown here applies only to

53 conditions giving the LR0 outcome curve, when the projections (and probabilities) are the same for LR and QM models.

Figure 3. The program window, showing typical outcome curves with different parameter set- tings. Settings for controls were as follows: A. The HV/VH photon source of 1,000 orthogonally correlated pairs at a fixed orientation (H is 0o). Malus’ law is implemented (Malus LR mode), polarizer 1 is fixed (at 0o), and polarizer 2 rotated at 50 random angles. Three coincidence count outcomes are shown (left panel), all of which follow the curves predicted from QM (LR0 curve): the default anti-correlations count (●); coincidence count (□); CHSH count (∆, right scale). The right panel shows the single counts at all four detectors, and their mean. The Gadgets (top section of window) show values from the last set; these were examined by clicking on the Sample # slider (values for the 30th pair shown). B. Photon source as in A, but with 5,000 pairs, each at random orientation. With polarizers in Malus LR mode, two successive Run1 cycles were made, one with polarizer 1 in ‘Fixed’ mode (●, anti-cor- relation; ○, CHSH count), and the second with both polarizer angles set randomly at 90 values (□, anti-correlation; ■, CHSH count). The correlations follow a sinusoidal curve, but with half the ampli- tude of QM expectations (LR2 curve). The right panel shows mean singles-counts as in A. C. HV/VH photon source rotated to λ = 45o, with polarizer1 also rotated to λ = 45o to give a common rotational frame. In a Run3 sequence, polarizer 2 was initially set to -45o to give an angle difference of -90o, and the angle was then rotated by 7.5o between successive runs, covering a range for angle difference from -90o to 90o. The outcome (the mean from 5 runs with 1,000 photon pairs at each an- gle difference; left panel; ●, anti-correlation; □, CHSH count, right scale) follows the LR0 curve. The same outcome would have been found for any value of λ, as long as both photon source and static po- larizer were rotate to the same value in a common rotational frame. The Analog implementation (right panel) shows calculated points and theoretical curves for Malus’ law yield differences in the

O 2 ordinary rays of the polarizers for the two orthogonal photons of a pair: IHV = cos σ (dark blue

O 2 E 2 points and curve), and IVH = –cos σ, (light blue); and complementary curves IHV = sin σ and

E 2 IVH = –sin σ, (red and yellow) for the extraordinary rays. With the Malus’ law yields normalized,

O application to photon pair gives, for particular setting, α, of the fixed polarizer Eα,β (LR) = 0.5{( IHV -

54 O E E 2 2 2 IVH ) - ( IHV - IVH )} = cos σ - sin σ = cos2σ, etc., the same outcome as QM. Since sin x = 1 - cos2x, the curve is scaled to 2cos2σ, offset by -1, as with conventional scoring. D. Random photon source, with discriminators set in Bell binary mode, and successive Run1 simula- tions (1,000 photon pairs) taken with polarizer1 angle fixed at 0o (closed symbols), or with both po- larizers at random (open symbols). Points are shown for HV/VH source (● or ○, anti-correlation; ■ or □, CHSH count), and for HH/VV source (♦ or ◊, CHSH count and Δ, anti-correlation). All fall on the diagonals expected from Bell’s zigzag (LR1 curve). The right-panel shows the singles-counts. E. HV/VH source, anti-correlation counts under the following rotations of polarizer 1 (P1) and pho- ton frames: ●, P1 0o, photon 0o; ■, P1 45o, photon 45o; □, P1 45o, photon 0o; ○, P1 0o, photon 45o; ◊, P1 30o, photon 0o; ♦, P1 0o, photon 30o; ∆, P1 30o, photon 30o. The right panel shows analog points and curves for the ordinary rays (see C) that illustrate why amplitude is lost on rotation of frames.

Figure 4. The Bell-type inequalities examined. LR0 and LR1 curves from anti-correlation (●, □, left scale, red) and CHSH counts (●, □, right scale, gray) for populations of 1,000 pairs of HV/VH photons, simulated through Run3 cycles, with photon source static and ‘Malus LR’ polarizer mode, (LR0, closed symbols), or photon source random and ‘Bell binary’ polarizer mode (LR1, open sym- bols). The magenta points and symbols show the difference between the LR0 and LR1 curves (in the spirit of the δ-function). The dashed lines show the intercepts at canonical values used to calculate the SBCHSH expectations. See text for discussion.

Figure 5. Simulation of objective local and quantum mechanical outcome curves. Successive Run1 simulations using HV/VH photon populations were performed as follows: ●, aligned frames at 0o (LR0 outcome); ○, polarizer 1 (P1) at -45o; ∆, P1 and photon frame both at -45o (LR0 outcome); ∆, P1 at -45o, QM mode; ■, stochastic photons, P1 at -45o (LR2 outcome); □, stochastic photons, QM mode. Note that at all these settings, the single-counts of the right panel showed the same pattern of behavior (yield of 0.5 at each detector).

Figure 6. Entropic penalty for stochastic orientation in an LR population. The sample enve- lope of the Malus’ law expectation curves in A, or the distribution of Malus’ law yield differ- ences revealed through the Analog option in B, provide a visualization of the penalty. The pho-

55 ton population was stochastic, in HV/VH state, and coincidences were detected used the anti-cor- relation count. A Run3 protocol using an average of 5, with polarizer 1 at 0o, generated both pan- els. A: Experimental points (●), and red curve show the LH2 outcome when the photon popu- lation is stochastic. The green lines are 125 theoretical expectation curves, calculated using 360 values of σ over the full circle, but normalized to the ±90o range shown. Each point was calcu- lated from 8 cross-products of Malus’ law yields at detectors Q, R, S, T (QS, RS, RT, and QT, for HV and VH configurations, appropriately weighted), generated using the random angle of orien- tation for the first pair of each population used at the particular setting of polarizer 1 (each line showing the curve expected if a population of photon pairs with this orientation was determined at each value for σ). The curves map out a sample envelope of values, equally distributed in phase space around the mean curve, so that the mean from the stochastic population of photon pairs is extracted from within such a range of values. B: The Malus’ law yield for each photon (from the population of 5,000 in the average of 5) in the ordinary and extraordinary rays of the two polarizers was calculated at each angle dif- ference. The differences in yield for a correlated pair measured in the ordinary (dark and light blue) and extraordinary (red, yellow) rays were then plotted for each hemisphere. Every one of the photon pairs gave a different yield, to give a range of values in a vertical bar at each angle difference, which fall either side of the theoretical curves. These follow a 푐표푠휑 (Malus’ law) distribution of values, where φ = θ – λ, θ is the orientation of the variable polarizer (with the fixed polarizer at 0o, θ = σ), and λ is the vector of the photon. The curve is centered at θ, and varies with λ. For a photon population at fixed orientation, all points calculated at a particular angle difference would overlap as a single point and follow the theoretical curves ±cos2σ (dark and light blue) or ±sin2σ (red, yellow). The points following the LR2 curve are the mean values from coincidences, derived with appropriate sign from elemental cross-products. In the anti-correlation count used here this is implicit, but, for example from the CHSH count (∑(푄푆 + 푅푆 + 푅푇 − 푄푇))/푛 (see Program Notes) it is explicit. All algorithms (with account taken of the Bell-state) follow the half-visibility curve, with envelopes to match. Note that all photons contribute to the outcome curve (left panel), but that cancellations lead to the analytical outcome, which segregates two equal contributions, both of which depend

56 on σ. This is shown in the left panel by the fractional amplitudes inside and outside the envelope, each of which at any value for σ, sum to half the total amplitude expected from the LR0 curve.

Figure 7. Comparison of a simulation of the Freedman-Clauser treatment (blue symbols) with the standard vOL simulation. The treatment underlying the δ-function of Freedman and Clauser22 leads to an OL outcome that matches the experimental result. Left panel: Symbols show means from measurement of 10,000 photon pairs when coincidences were counted at 100 angle differences using the FC δ-count protocol with different sources. Points show: static HH/VV source (●, the LR0 curve at the half-scale expected from normalization and an equal mix of H and V); random HH/VV source (♦, the LR2 curve at half-scale, which shows half the amplitude of the LR0 curve, as expected from an isotropic source); and PDC EO source with V photons in channel 2 rotated to H by HWP2 at 45o (■, this generates a fully polarized source with both channels in H orientation). This last source would represent the NL outcome expect after alignment of a stochastic source with polarizer 1 set at 0o (see text). The red symbols show the outcome from the standard coincidence count at the same scale (●, static HH/VV source; ♦, random source). Right panel: mean singles-counts show the HH/VV counts, all distributed with values close to 0.5 as expected from an isotropic source. However, with the PDC EO source, where both signal (ordinary) and idler (extraordinary) channels are polarized. Then, counts with values either at 1 and 0 (at the station with the fixed polarizer), or as complemen- tary cosine and sine curves (the polarization detected at the other station). For each value of σ, the four mean counts at any angle difference sum to give the two photons of a pair.

Figure 8. Simulation of early PDC-based experiments claimed as supporting non-locality17,41,80. A. Curve 1 shows successive runs as follows: ●, LR0 curve (cf. Figs. 1, 2, no ‘state prepara- tion’); ○, HWP0 set at 45o (same curve). Curves 2-4 show the effects of setting “the analyzer in beam 1 at 45o”: Curve 2 (□) was obtained by setting HWP1 at ±45o in front of static polarizer 1 at 0o (one literal interpretation of the text in Kwiat et al.41 ); for curves 3, successive runs were taken first with HWP1 rotated by 22.5o with polarizer 1 fixed at 0o (Δ), then with the HWP1 at 0o, and polarizer 1 at 45o (■, which is equivalent, both showing zero visibility). Curve 4 (●) shows the outcome on rotation both of HWP1 by 22.5o and of polarizer 1 by 45o. Curves 2 and 4 show the same full visibility as curve 1, but curve 2 is phase-shifted by 90o, and curve 4 by 45o. This is the behavior shown in Fig. 7 of Fiorentino et al.80, and interpreted as demonstrating the QM expectation.

57 B. With polarizer 1 fixed at 0o, HWP2 was introduced in beam 2, in which the setting of po- larizer 2 was separately varied over the full range. Successive curves show the outcome as HWP2 was rotated by increments of 15o over the range θ = ±45o. All curves show full visibility, but are phase displaced by 2θ, the angle through which the beam was rotated; the behavior is fully OL com- patible. This outcome is, in effect, the same as that reported by Weihs et al.17 using EOMs to rotate the beams. The outcome is also similar to that found experimentally in80,86, and claimed there to show the full rotational invariance expected from NL predicates. Although the outcome here has the fixed polarizer at 0o, and has both the offset and analysis rotations in the variable channel, it demonstrates that the configuration (which matches in essentials those in80,86) can be used to generate outcome curves that appear to show full-visibility rotational invariance on rotation of one polarizer.

58 Figures

Figure 1

59

Figure 2

60

Figure 3A

61

Figure 3B

62

Figure 3C

63

Figure 3D

64

Figure 3E

65

Figure 4

66

Fig. 5

67

Fig. 6

68

Fig. 7

69

Fig. 8

70 Bibliography

1. Einstein, A.; Podolsky, B.; Rosen, N. Can Quantum-Mechanical Description of Physical Reality Be Considered Complete? Physical Review 1935, 47, 777-780. 2. Fine, A. The Natural Ontological Attitude. In The Shaky Game: Einstein, Realism and the Quantum Theory. Fine, A., Ed.; University of Chicago Press: Chicago, 1986, p p. 116. 3. Isaacson, W. Einstein: His Life and Universe. Simon & Schuster Paperbacks: New York, 2008. 4. Gilder, L. The Age of Entanglement: when quantum physics was reborn. Alfred A. Knopf: New York, 2008. 5. Fine, A. The Einstein-Podolsky-Rosen Argument in Quantum Theory. In The Stanford Encyclopedia of Philosophy (Edward N. Zalta, ed.); Metaphysics Research Lab, Stanford University: Stanford, 2004. 6. Howard, D. Was Einstein really a realist? Perspectives in Physics 1993, 1, 204-251. 7. Howard, D. A. as a Philosopher of Science. Physics Today 2005, 58, 34- 40. 8. Bohr, N. On the notions of causality and complementarity. Dialectica 1948, 2, 312-319. 9. Bohm, D.; Aharonov, Y. Discussion of experimental proof for the paradox of Einstein, Rosen, and Podolsky. Physical Review 1957, 108, 1070-1076. 10. Einstein, A. Einstein: His Life and Universe. Isaacson, W., Ed.; AEA 22-174, 1950. 11. Stapp, H. P. Bell's theorem and world process. Nuovo Cimento B 1975, 29, 270-276. 12. Leggett, A. J. Bell's Theorem. In Compendium of Quantum Physics: Concepts, Experiments, History and Philosophy. Greenberger, D.; Hentschel, K.; Weinert, F., Eds.; Springer-Verlag Berlin, Heidelberg, 2009, p 24-31. 13. Bell, J. S. On the Einstein-Podolsky-Rosen paradox. Physics Today 1964, 1, 195-200. 14. Bell, J. S. On the problem of hidden variables in quantum mechanics. Rev. Mod. Phys. 1966, 38, 312-319. 15. Clauser, J. F.; Horne, M. A.; Shimony, A.; Holt, R. A. Proposed experiment to test local hidden-variable theories. Physical Review Letters 1969, 23, 880-884. 16. Aspect, A. Bell's inequality test: more ideal than ever. Nature 1999, 398, 189-190. 17. Weihs, G.; Jennewein, T.; Simon, C.; Weinfurter, H.; Zeilinger, A. Violation of Bell's inequality under strict Einstein locality conditions. Phys. Rev. Lett. 1998, 81, 5039-5044. 18. Zeilinger, A. Experiment and the foundations of quantum physics. Reviews of 1999, 71, S288-S297. 19. Shimony, A. Bell's Theorem. In The Stanford Encyclopedia of Philosophy (Winter 2013 Edition). Zalta, E. N., Ed.; Metaphysics Research Lab, Stanford University: Stanford, 2009. 20. Bell, J. S. Einstein-Podolsky-Rosen experiments. In Proceedings of the symposium on Frontier problems in high-energy physics; Annali Della Scola Normale di Pisa: Pisa, 1976, p 33- 45. 21. Clauser, J. F.; Horne, M. A. Experimental consequences of objective local theories. Physical Rev. D. 1974, 10, 526-535. 22. Freedman, S. J.; Clauser, J. F. Experimental test of local hidden-variable theories. Physical Review Letters 1972, 24, 938-941. 23. Lochhead, J. Einstein's Quantum Riddle. In NOVA; PBS, 2019.

71 24. Fine, A. Science Made Up: Constructivist Sociology of Scientific Knowledge. In The Disunity of Science: Boundaries, Contexts, and Power. Galison, P.; Stump, D., Eds.; Stanford University Press, 1996, p 231-254. 25. Bub, J. Quantum Entanglement and Information. In The Stanford Encyclopedia of Philosophy (Winter 2008 Edition), Edward N. Zalta (ed.); Metaphysics Research Lab, Stanford University: Stanford, 2008. 26. Faye, J. Copenhagen Interpretation of Quantum Mechanics. In The Stanford Encyclopedia of Philosophy (Zalta, E.N., ed); Metaphysics Research Lab, Stanford University: Stanford, 2008. 27. Heisenberg, W. Physics and Philosophy. Harper and Row: New York, 1958. 28. Herbert, N. Quantum Reality: Beyond the New Physics Anchor Books: New York, 1987. 29. Mermin, N. D. Hidden variables and the two theorems of John Bell. Reviews of Modern Physics 1993, 65, 803-815. 30. Kocsis, S.; Braverman, B.; Ravets, S.; Stevens, M. J.; Mirin, R. P.; Shalm, L. K.; Steinberg, A. M. Observing the Average Trajectories of Single Photons in a Two-Slit Interferometer. Science 2011, 332, 1170-1173. 31. Atmanspacher, H. Quantum Approaches to Consciousness. In The Stanford Encyclopedia of Philosophy (Summer Edition, Edward N. Zalta, ed.). Zalta, E. N., Ed., 2011. 32. Bernstein, J.; Holt, J. Spooky Physics Up Close: An Exchange. In The New York Review of Books, 12/08 issue; NYREV, Inc.: New York, NY, 2016. 33. Gröblacher, S.; Paterek, T.; Kaltenbaek, R.; Brukner, Č.; Żukowski, M.; Aspelmeyer, M.; Zeilinger, A. An experimental test of non-local realism. Nature 2007, 446, 871-875. 34. Zeilinger, A. Dance of the Photons: From Einstein to Quantum Teleportation. Farrar, Straus and Giroux: New York, N.Y., 2010. 35. Bell, J. S. Introduction to the hidden variables question. In Proceeding of the International School of Physics 'Enrico Fermi': Foundations of Quantum Mechanics d'Espagnat, B., Ed.; Academic Press: New York, 1971, p 171-181. 36. Fry, E. S.; Thompson, R. C. Experimental test of local hidden-variables theories. Rev. Mod. Phys. 1976, 37, 465-468. 37. Aspect, A.; Grangier, P.; Roger, G. Experimental tests of realistic local theories via Bell's theorem. Phy. Rev. Lett. 1981, 47, 460-463. 38. Aspect, A.; Dalibard, J.; Roger, G. Experimental test of Bell's inequalities using time- varying analyzers. Phys. Rev. Lett. 1982, 49, 1802-1807. 39. Kocher, C. A.; Commins, E. D. Polarization correlation of photons emitted in an atomic cascade. Phys. Rev. Lett. 1967, 18, 575-577. 40. Aspect, A.; Grangier, P.; Roger, G. Experimental Realization of Einstein-Podolsky- Rosen-Bohm Gedankenexperiment: A New Violation of Bell's Inequalities. Phys. Rev. Lett. 1982, 49, 91-94. 41. Kwiat, P. G.; Mattle, K.; Weinfurter, H.; Zeilinger, A.; Sergienko, A. V.; Shih, Y. New high-intensity source of polarization-entangled photon pairs. Phys. Rev. Lett. 1995, 75, 4337- 4341. 42. Kwiat, P. G.; Waks, E.; White, A. G.; Appelbaum, I.; Eberhard, P. H. Ultrabright source of polarization-entangled photons. Phys. Rev. A. 1999, 60, R773-R776. 43. Tittel, W.; Brendel, J.; Gisin, B.; Herzog, T.; Zbinden, H.; Gisin, N. Experimental demonstration of quantum-correlations over more than 10 km. Phys. Rev. A. 1998, 57, 3229- 3232.

72 44. Salart, D.; Baas, A.; Branciard, C.; Gisin, N.; Zbinden, H. Testing the speed of "spooky action at a distance". Nature 2008, 454, 861-864. 45. B. Wittmann, S. R., F. Steinlechner, N. K. Langford, N. Brunner, H. Wiseman, R. Ursin, A. Zeilinger. Loophole-free Einstein–Podolsky–Rosen experiment via quantum steering. New Journal of Physics 2012, 14, 053030. 46. Ma, X.-S.; Herbst, T.; Scheidl, T.; Wang, D.; Kropatschek, S.; Naylor, W.; Wittmann, B.; Mech, A.; Kofler, J.; Anisimova, E.; Makarov, V.; Jennewein, T.; Ursin, R.; A., Z. Quantum teleportation over 143 kilometres using active feed-forward. Nature 2012, 489, 269-273. 47. Giustina, M.; Mech, A.; Ramelow, S.; Wittmann, B.; Kofler, J.; Beyer, J.; Lita, A.; Calkins, B.; Gerrits, T.; Nam, S. W.; Ursin, R.; Zeilinger, A. Bell violation using entangled photons without the fair-sampling assumption. Nature 2013, 497, 227-230. 48. Peres, A. Quantum Theory: Concepts and Methods. Kluwer Academic Publ.: New York, 2002. 49. Garg, A.; Leggett, A. J. Comment on "Bell's Theorem: Does the Clauser-Horne inequality hold for all local theories?". Phys. Rev. Lett. 1984, 53, 1019-1020. 50. Holt, R. A. Atomic Cascade Experiments. In Physics; Harvard University: Cambridge. Mass., 1973. 51. Gröblacher, S.; Paterek, T.; Kaltenbaek, R.; Brukner, Č.; Żukowski, M.; Aspelmeyer, M.; Zeilinger, A. An experimental test of non-local realism. Nature 2007, 446, 871-875. 52. Neumann, P.; Mizuochi, N.; Rempp, F.; Hemmer, P.; Watanabe, H.; Yamasaki, S.; Jacques, V.; Gaebel, T.; Jelezko, F.; Wrachtrup, J. Multipartite Entanglement Among Single Spins in Diamond. Science 2008, 320, 1326-1328. 53. Aspect, A.; Grangier, P.; Roger, G. Experimental Tests of Realistic Local Theories via Bell's Theorem. Physical Review Letters 1981, 47, 460-463. 54. Altepeter, J. B.; Jeffrey, E. R.; Kwiat, P. G. Phase-compensated ultra-bright source of entangled photons. Express 2005, 13, 8951-8959. 55. Christensen, B. G.; McCusker, K. T.; Altepeter, J. B.; Calkins, B.; Gerrits, T.; Lita, A. E.; Miller, A.; Shalm, L. K.; Zhang, Y.; S. W. Nam; Brunner, N.; Lim, C. C. W.; Gisin, N.; Kwiat, P. G. Detection-Loophole-Free Test of Quantum Nonlocality, and Applications. Phy. Rev. Lett. 2013, 111, 130406. 56. Shalm, L. K.; Meyer-Scott, E.; Christensen, B. G.; Bierhorst, P.; Wayne, M. A.; Stevens, M. J.; Gerrits, T.; Glancy, S.; Hamel, D. R.; Allman, M. S.; Coakley, K. J.; Dyer, S. D.; Hodge, C.; Lita, A. E.; Verma, V. B.; Lambrocco, C.; Tortorici, E.; Migdall, A. L.; Zhang, Y.; Kumor, D. R.; Farr, W. H.; Marsili, F.; Shaw, M. D.; Stern, J. A.; Abellán, C.; Amaya, W.; Pruneri, V.; Jennewein, T.; Mitchell, M. W.; Kwiat, P. G.; Bienfang, J. C.; Mirin, R. P.; Knill, E.; Nam, S. W. A strong loophole-free test of local realism. In arXiv; Cornell University Library: Ithaca, N.Y., 2015, p http://arxiv.org/abs/1511.03189. 57. Bell, J. S. Locality in quantum mechanics: reply to critics. Epistemological Letters 1975, Nov., 2-6. 58. Yin, J.; Cao, Y.; Li, Y.-H.; Liao, S.-K.; Zhang, L.; Ren, J.-G.; Cai, W.-Q.; Liu, W.-Y.; Li, B.; Dai, H.; Li, G.-B.; Lu, Q.-M.; Gong, Y.-H.; Xu, Y.; Li, S.-L.; Li, F.-Z.; Yin, Y.-Y.; Jiang, Z.- Q.; Li, M.; Jia, J.-J.; Ren, G.; He, D.; Zhou, Y.-L.; Zhang, X.-X.; Wang, N.; Chang, X.; Zhu, Z.- C.; Liu, N.-L.; Chen, Y.-A.; Lu, C.-Y.; Shu, R.; Peng, C.-Z.; Wang, J.-Y.; Pan, J.-W. Satellite- based entanglement distribution over 1200 kilometers. Science 2017, 356, 1140-1144. 59. Handsteiner, J.; Friedman, A. S.; Rauch, D.; Gallicchio, J.; Liu, B.; Hosp, H.; Kofler, J.; Bricher, D.; Fink, M.; Leung, C.; Mark, A.; Nguyen, H. T.; Sanders, I.; Steinlechner, F.; Ursin,

73 R.; Wengerowsky, S.; Guth, A. H.; Kaiser, D. I.; Scheidl, T.; Zeilinger, A. Cosmic Bell Test: Measurement Settings from Milky Way Stars. Physical Review Letters 2017, 118, 060401. 60. Hensen, B.; Bernien, H.; Dréau, A. E.; Reiserer, A.; Kalb, N.; Blok, M. S.; Ruitenberg, J.; Vermeulen, R. F. L.; Schouten, R. N.; Abellán, C.; Amaya, W.; Pruneri, V.; Mitchell, M. W.; Markham, M.; Twitchen, D. J.; Elkouss, D.; Wehner, S.; Taminiau, T. H.; Hanson, R. Loophole- free Bell inequality violation using electron spins separated by 1.3 kilometres. Nature 2015, 526 682-686. 61. Mermin, N. D. Is the moon there when nobody looks? Reality and the quantum theory Physics Today 1985, 38, 38-47. 62. Thompson, C. H. Rotational Invariance, Phase Relationships and the Quantum Entanglement Illusion. In arXiv; Cornell University Library: Ithaca, N.Y., 1999, p 9912082. 63. De Raedt, K.; De Raedt, H.; Michielsen, K. Deterministic event-based simulation of quantum phenomena. Computer Physics Communications 2005, 171, 19-39. 64. Ladyman, J.; Linnebo, Ø.; Bigaj, T. Entanglement and non-factorizability. Studies in History and Philosophy of Modern Physics 2013, 44, 215-221. 65. Angelidis, T. D. Bell's theorem: does the Clauser-Horne inequality hold for all local theories? Phys. Rev. Lett. 1983, 51, 1819-1822. 66. Angelidis, T. D. On the problem of a local extension of the quantum formalism. J. Math. Phys. 1993, 34, 1635-1653. 67. Wesley, J. P. Experimental results of Aspect et al. confirm classical local causality. Physics Essays 1994, 7, 240-242. 68. Thompson, C. H. The Chaotic Ball: An Intuitive Analogy for EPR Experiments. Found.Phys.Lett. 1996, 9, 357-382. 69. Hess, K.; Philipp, W. Breakdown of Bell's theorem for certain objective local parameter spaces. Proc. Natl. Acad. Sci. (U.S.A) 2004, 101 1799-1805. 70. Christian, J. Disproof of Bell’s Theorem by Clifford Algebra Valued Local Variables In arXiv; Cornwell University Library: Ithaca, N.Y., 2007, p 0703179. 71. De Raedt, K.; De Raedt, H.; Michielsen, K. A computer program to simulate Einstein– Podolsky–Rosen–Bohm experiments with photons. Computer Physics Communications 2007, 176, 642–651. 72. De Zela, F. A non-local hidden-variable model that violates Leggett-type inequalities. J. Physics A 2008, 41, 50531-50540. 73. Thompson, C. H. Subtraction of "accidentals'' and the validity of Bell tests. In arXiv, 1999, p 9903066. 74. Popper, K. R. Quantum Theory and the Schism in Physics. Hutchinson: London, 1982, p 27-30. 75. Bell, J. S. Bertlmann's socks and the nature of reality. Journal de Physique 1981, 42, C2 41-61. 76. Jaynes, E. T. Clearing up the mysteries - the original goal. In Maximum Entropy and Bayesian Methoods. Skilling, J., Ed.; Kluwer Academic Publishers: Dordrecht, Holland, 1989, p 1-27. 77. Christian, J. Disproof of Bell’s Theorem: Reply to Critics. In arXiv; Cornell University Library: Ithaca, N.Y., 2008, p 0703244. 78. Hiley, B. J.; Callaghan, R. E. The Clifford Algebra approach to Quantum Mechanics A: The Schrödinger and Pauli Particles. In arXiv, 2010.

74 79. Hiley, B. J.; Callaghan, R. E. The Clifford Algebra Approach to Quantum Mechanics B: The Dirac Particle and its relation to the Bohm Approach. In arXiv, 2010, p 4033. 80. Fiorentino, M.; Kuklewicz, C.; Wong, F. Source of polarization entanglement in a single periodically poled KTiOPO4 crystal with overlapping emission cones. Optical Express 2005, 13, 127-135. 81. Bell, J. S. Speakable and Unspeakable in Quantum Mechanics. Cambridge University Press: Cambridge U.K., 1987. 82. Einstein, A. Letter to Erwin Schrödinger. In Letter to Erwin Schrödinger, Dec. 1950, quoted in "Einstein: His Life and Universe" by Isaacson, W.; AEA 22-174, 1950. 83. Albert, D. Z. Quantum Mechanics and Experience. Harvard University Press: Harvard, MA, 1992. 84. Kwiat, P. G.; Hardy, L. The mystery of the quantum cakes. Am. J. Physics 1999, 68 33- 36. 85. Einstein, A. Physik und realität. J. Franklin Institute 1936, 221, 313-448. 86. Kim, T.; Fiorentino, M.; Wong, F. N. C. Phase-stable source of polarization-entangled photons using a polarization Sagnac interferometer. Phys. Rev. A 2006, 73, 012316. 87. Nagata, K.; Laskowski, W.; Wieśniak, M.; Żukowski, M. Rotational invariance as an additional constraint on local realism. Phys. Rev. Lett. 2004, 93, 230403 230401-230404. 88. von Neumann, J. Mathematical Foundations of Quantum Mechanics (English translation, 1955). Princeton University Press: Princeton, N.J., 1932. 89. Mermin, N. D.; Schack, R. Homer Nodded: Von Neumann’s Surprising Oversight. Foundations of Physics 2018, 48, 1007-1020. 90. Bialynicki-Birula, I.; Bialynicki-Birula, Z. Uncertainty Relation for Photons. Phys.L Rev. Lett. 2012, 108, 140401-140405. 91. Bell, J. S. Six possible worlds of quantum mechanics. In: Proceedings of the Nobel Symposium : Possible Worlds in Arts and Sciences, Stockholm. 1986; pp 11-15. 92. Dirac, P. A. M. On the Theory of Quantum Mechanics. Proceedings of the Royal Society A. 1926, 112 (762), 661-677. 93. Dirac, P. A. M. The Quantum Theory of the Electron. Proceedings of the Royal Society of London A. 1928, 117 (778), 610-624. 94. Dirac, P. A. M. The Principles of Quantum Mechanics (4th. Edition). Oxford University Press: Oxford, 1958. 95. Wikipedia-entry. Interpretation of quantum mechanics, 2010. 96. Migdall, A. Polarization directions of noncollinear phase-matched optical parametric downconversion output. J. Opt. Soc. Am. B 1997, 14, 1093-1098. 97. Kwiat, P. G.; Eberhard, P. H.; Steinberg, A. M.; Chiao, R. Y. Proposal for a loophole-free Bell inequality experiment. Physical Rev. A 1994, 49, 3209-3220. 98. Gerlach, W.; Stern, O. Das magnetische Moment des Silberatoms. Zeitschrift für Physik 1922, 9, 353-355. 99. Feynman, R. The Development of the Space-Time View of Quantum Electrodynamics. In: Nobel Lectures, Stockholm, Sweden. Nobel Foundation, 1965. 100. Jameson, D. M. Gregorio Weber, 1916–1997: A Fluorescent Lifetime. Biophys. J. 1998, 75, 419-421. 101. Gratton, E.; Barbieri, B. B. Multifrequency phase fluorometry using pulsed sources: theory and applications. Spectroscopy 1986, 1, 28-36.

75 102. Popper, K. R. Objective Knowledge: An Evolutionary Approach. Clarendon Press: Oxford, 1972. 103. Ursin, R.; Tiefenbacher, F.; Schmitt-Manderbach, T.; Weier, H.; Scheidl, T.; Lindenthal, M.; Blauensteiner, B.; Jennewein, T.; Perdigues, J.; Trojek, P.; Ömer, B.; Fürst, M.; Meyenburg, M.; Rarity, J.; Sodnik, Z.; Barbieri, C.; Weinfurter, H.; Zeilinger, A. Entanglement-based quantum communication over 144 km. Nature Physics 2007, 3, 481-486. 104. Scheidl, T.; Ursin, R.; Kofler, J.; Ramelow, S.; Ma, X.-S.; Herbst, T.; Ratschbacher, L.; Fedrizzi, A.; Langford, N. K.; Jennewein, T.; Zeilinger, A. Violation of local realism with freedom of choice. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 19708-19713. 105. Wittmann, S., B. ; Steinlechner, F.; Langford, N. K.; Brunner, N.; Wiseman, H.; Ursin, R.; Zeilinger, A. Loophole-free Einstein–Podolsky–Rosen experiment via quantum steering. New Journal of Physics 2012, 14, 053030. 106. Crofts, A. R. Life, information, entropy, and time. Complexity 2007, 13, 14-50. 107. Merali, Z. Quantum mechanics braces for the ultimate test. Science 2011, 331, 1380- 1382. 108. Eberhard, P. Bell's theorem and the different concepts of locality. Nuovo Cimento 1978, 46B, 392-419. 109. Penrose, R. The Emperor's New Mind : concerning computers, minds, and the laws of physics. Penguin Books: New York, N.Y., 1991. 110. Davies, P.; Gregersen, N. H. Information and the nature of reality. Cambridge University Press: Cambridge, UK, 2010. 111. Kaiser, D. How the Hippies Saved Physics: Science, Counterculture, and the Quantum Revival. W.W. Norton & Co., Inc.: New York, N.Y., 2011. 112. Peres, A. Einstein, Podolsky, Rosen, and Shannon In arXiv:quant-ph/0310010 v1, 2003. 113. Zeilinger, A. On the Interpretation and Philosophical Foundation of Quantum Mechanics. In Die Wiissenschaft an den Grenzen der Erde; Michael Imhof, 2008. 114. Brown, D. Angels and Demons. Pocket Books (Simon & Schuster): New York, N.Y., 2000. 115. Brown, D. The Lost Symbol. Random House: New York, 2009. 116. Wikipedia-contributors. Polarization (waves). In Wikipedia, The Free Encyclopedia; Wikipedia, The Free Encyclopedia: https://en.wikipedia.org/w/index.php?title=Polarization_(waves)&oldid=912670511, 2019. 117. Feynman, R. P. QED. Princeton University Press: Princeton, New Jersey, 1985. 118. Feynman, R. P. Space-time approach to non-relativistic quantum Mechanics. Rev. Mod. Physics 1948, 20, 367-387. 119. Wikipedia_contributors. Vacuum state In Wikipedia, The Free Encyclopedia https://en.wikipedia.org/w/index.php?title=Vacuum_state&oldid=921400081, 2019. 120. Wikipedia_contributors. Vacuum polarization. In Wikipedia, The Free Encyclopedia, : https://en.wikipedia.org/w/index.php?title=Vacuum_polarization&oldid=931343429, 2019. 121. Shao, Z. Refractive indices for extraordinary waves in uniaxial crystals. Physical Review E 1995, 52, 1043-1048. 122. Bialynicki-Birula, I. Photon Wave Function. In Progress in Optics XXXVI. Wolf, E., Ed.; Elsevier: Amsterdam, 1996, p p. 245-294. 123. Zalesny, J. Derivation of the Bialynicki-Birula Photon WaveFunction From Three- Component Spinors. International Journal of Theoretical Physics 2004, 43, 2093-2100.

76 124. Maxell. What is the relation between electromagnetic wave and photon? In Physics Stack Exchange; https://physics.stackexchange.com/q/438929, 2018. 125. Wódkiewicz, K. Propensity and probability in quantum mechanics. Physics Letters A 1986, 115, 304-306.

77